ArticlePDF AvailableLiterature Review

COVID-19: systemic pathology and its implications for therapy

Authors:

Abstract and Figures

Responding to the coronavirus disease 2019 (COVID-19) pandemic has been an unexpected and unprecedented global challenge for humanity in this century. During this crisis, specialists from the laboratories and frontline clinical personnel have made great efforts to prevent and treat COVID-19 by revealing the molecular biological characteristics and epidemic characteristics of the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). Currently, SARS-CoV-2 has severe consequences for public health, including human respiratory system, immune system, blood circulation system, nervous system, motor system, urinary system, reproductive system and digestive system. In the review, we summarize the physiological and pathological damage of SARS-CoV-2 to these systems and its molecular mechanisms followed by clinical manifestation. Concurrently, the prevention and treatment strategies of COVID-19 will be discussed in preclinical and clinical studies. With constantly unfolding and expanding scientific understanding about COVID-19, the updated information can help applied researchers understand the disease to build potential antiviral drugs or vaccines, and formulate creative therapeutic ideas for combating COVID-19 at speed.
Content may be subject to copyright.
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
386
International Journal of Biological Sciences
2022; 18(1): 386-408. doi: 10.7150/ijbs.65911
Review
COVID-19: systemic pathology and its implications for
therapy
Qi Shen1,2, Jie Li1,2, Zhan Zhang3,4, Shuang Guo5, Qiuhong Wang6, Xiaorui An1,2, Haocai Chang1,2
1. MOE Key Laboratory of Laser Life Science & Institute of Laser Life Science, College of Biophotonics, South China Normal University, Guangzhou 510631,
China.
2. Guangdong Provincial Key Laboratory of Laser Life Science, College of Biophotonics, South China Normal University, Guangzhou 510631, China.
3. Department of Neurology, Sun Yat-sen Memorial Hospital, Sun Yat-sen University, Guangzhou 510120, China.
4. Guangdong Province Key Laboratory of Brain Function and Disease, Zhongshan School of Medicine, Sun Yat-sen University, Guangzhou 510120, China.
5. Dermatology Hospital, Southern Medical University, Guangzhou 510091, China.
6. Qilu Cell Therapy Technology Co., Ltd, Jinan 250000, China.
Corresponding author: changhc@scnu.edu.cn, College of Biophotonics, South China Normal University, 55 West Zhongshan Avenue, Tianhe District,
Guangzhou 510631, China. Tel: +86-20-85210089; Fax: +86-20-85216052
© The author(s). This is an open access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/).
See http://ivyspring.com/terms for full terms and conditions.
Received: 2021.08.09; Accepted: 2021.11.04; Published: 2022.01.01
Abstract
Responding to the coronavirus disease 2019 (COVID-19) pandemic has been an unexpected and
unprecedented global challenge for humanity in this century. During this crisis, specialists from the
laboratories and frontline clinical personnel have made great efforts to prevent and treat COVID-19 by
revealing the molecular biological characteristics and epidemic characteristics of the severe acute
respiratory syndrome coronavirus 2 (SARS-CoV-2). Currently, SARS-CoV-2 has severe consequences
for public health, including human respiratory system, immune system, blood circulation system, nervous
system, motor system, urinary system, reproductive system and digestive system. In the review, we
summarize the physiological and pathological damage of SARS-CoV-2 to these systems and its molecular
mechanisms followed by clinical manifestation. Concurrently, the prevention and treatment strategies of
COVID-19 will be discussed in preclinical and clinical studies. With constantly unfolding and expanding
scientific understanding about COVID-19, the updated information can help applied researchers
understand the disease to build potential antiviral drugs or vaccines, and formulate creative therapeutic
ideas for combating COVID-19 at speed.
Key words: SARS-CoV-2, immune system, nervous system, reproductive system, motor system, immunotherapy
1. Introduction
Coronavirus disease 2019 (COVID-19) with acute
respiratory disease and potentially severe pneumonia
has resulted in unimaginable consequence to public
health and loss of human life due to severe acute
respiratory syndrome coronavirus 2 (SARS-CoV-2).
As the end of 5 November 2021, there have been
248,467,363 confirmed cases of COVID-19, including
5,027,183 deaths (https://covid19.who.int/). The
focus of global efforts is to develop safe and effective
vaccines against COVID-19, and some vaccines are
provisionally licensed for COVID-19 prevention.
However, the population of COVID-19 patients is still
surging in the case of public vaccination.
As reported, SARS-CoV-2 is a single-stranded
positive-sense RNA virus, whose genome size varies
from 29.8 ~ 29.9 kb [1], and whose genome encodes 4
structural proteins (spike surface glycoprotein (S),
membrane protein, envelope protein and
nucleocapsid protein), 16 non-structural proteins
(NSP1-16) and some accessory proteins (ORF3a,
ORF3b, ORF6, ORF7a, ORF7b, ORF8, ORF9b and
ORF10) [2-4] (Fig. 1). With unremitting efforts of
scientific researchers around the world, the structure
and function of these proteins have been more
understood. However, effective treatment including
drug or vaccine is still not available for COVID-19, the
study on SARS-CoV-2 still needs to be explored. So
far, few studies have systematically investigated the
Ivyspring
International Publisher
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
387
physiological and pathological features of
SARS-CoV-2 from a multisystem perspective (Fig. 1).
In this review, the information about the
characteristics, damage and treatment of SARS-CoV-2
is summarized and updated by our limited
understanding. Importantly, we have tried to give a
full insight into physiological effects and action
mechanisms of SARS-CoV-2 in order to find potential
therapeutic strategies or control measures in this way
to combat the ongoing infection.
2. Affect and Effect of COVID-19
2.1. COVID-19 on respiratory system
Obviously, respiratory system is the first to bear
the brunt of SARS-CoV-2, which as a danger signaling
is sent to immune system. As a receptor of S surface
protein on SARS-CoV-2, angiotensin converting
enzyme II (ACE2) is mainly expressed in type-2
alveolar (AT2) cells (83% in average) [5] and other
certain cells such as epithelial cells, which exists
widely in various tissues and organs including oral
mucosa [6], airway [7], lung [8], colon [9], kidney [10]
and prostate [11]. Depending on the context, lungs are
directly under the strongest attack to cause lung
inflammation and functional injury. As a result,
pulmonary fibrosis occurs frequently in the infected
patients, along with dry cough, dyspnea and fatigue
symptoms.
Pulmonary fibrosis is characterized by the
excessive deposition of extracellular matrix
components including collagen and fibronectin,
mainly originating from the persistence of fibroblast
proliferation and activation. A molecular explanation
for pulmonary fibrosis is that fibroblast growth factor
(FGF) and transforming growth factor (TGF), as
inducers of collagen and fibronectin [12, 13], are both
high levels in COVID-19 patients [14, 15]. As reported,
other upregulated pro-inflammatory cytokines, such
as platelet-derived growth factor (PDGF), vascular
endothelial growth factor (VEGF), tumor necrosis
factor α (TNFα) and interleukin-6 (IL-6), also involved
in the process [16]. Another important aspect is that
ongoing AT2 epithelial cell injury can recruit
fibroblasts to fibrotic loci, and then fibroblasts
differentiate into myofibroblasts to produce
extracellular matrix proteins [16, 17].
2.2. COVID-19 on immune system
An effective immune response against
SARS-CoV-2 requires two arms of the immune
system, the innate immune system and the adaptive
immune system. The innate immune system, as the
first line of defense of the immune system, is
responsible for rapidly recognizing the infection and
triggering alarms, in which a range of innate immune
cells are involved. On the other hand, the adaptive
immune system is essential for controlling and
clearing the viral infection, the three fundamental
components of which (CD4+ T cells, CD8+ T cells, B
cells) cooperate with each other to regulate the
antigen-specific immune responses. Therefore,
understanding the interaction between SARS-CoV-2
and the immune system is of great importance for
pathogenesis, disease progression, treatment
strategies, and vaccine design of COVID-19.
Figure 1. Structure of SARS-CoV-2 and its infection on tissues and organs in eight major systems of human organism.
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
388
Figure 2. The immunopathology of SARS-CoV-2. SARS-CoV-2 reduced the number of monocytes, NK cells, DCs, CD4+ T cells, CD8+ T cells and B cells. Conversely, the virus
increased the number of neutrophils, mast cells, macrophages, memory CD4+ T cells, memory CD8+ T cells and memory B cells to some extent, and triggered complement
system responses, then the host produced a strong and harmful cytokine storm, and a weak and favorable antibody response.
2.2.1. Innate immune response
The innate immune response correlates with the
severity of COVID-19, which has been proved by a
range of studies [18, 19]. The innate immune system
against SARS-CoV-2 infection has three main
pathways: (1) limiting virus replication in infected
cells; (2) development of antiviral state in the infection
site, including recruitment of various innate immune
cells; (3) initiating the adaptive immunity [20]. All of
these pathways require the involvement of multiple
types of innate immune cells, of which the more
common cells are granulocytes, monocytes,
macrophages and natural killer (NK) cells, but there
are also many other immune cells including different
dendritic cells (DCs), innate lymphoid cells, and mast
cells.
2.2.1.1. Granulocytes and monocytes
A few studies have revealed that increases in
monocytes, neutrophils and eosinophils correlate
with the severity of disease in COVID-19 patients [21]
(Fig. 2). Liu found that the neutrophil-to-lymphocyte
ratio (NLR) could be served as a risk factor for
early-stage prediction of COVID-19 patients, whose
age is over 50 years old and NLR is more than 3.13 are
more likely predicted to develop critical illness [22]. A
prominent higher proportion of neutrophils and
activated mast cells was also observed in the
bronchoalveolar lavage fluid (BALF) of COVID-19
patients [18] (Fig. 2). The fact that chemoattractants
for neutrophils were upregulated during COVID-19
[19, 23, 24] suggested that granulocytes may
significantly contribute to pathogenesis.
CD16+ monocytes were depleted in COVID-19
patients and were remodeled with a phenotypic shift
toward CD14+ monocytes [24]. Inflammatory
HLA-DRhi CD11chi CD14+ monocytes can be
considered as a hallmark of mild COVID-19 patients
[25]. Moreover, decrease of CD14lo CD16hi monocytes
and the appearance of dysfunctional monocytes with
low expression of human leukocyte antigen class DR
(HLA-DR) and high expression of alarming S100 were
related to the severe COVID-19 patients [24, 25].
Interestingly, Guo et al. found a unique subpopulation
of monocytes, with high expression of inflammatory
genes, specifically present in the severe COVID-19
cases. These inflammatory genes were potentially
regulated by the transcription factors ATF3, NFIL3,
and HIVEP2 [26]. And COVID-19 patients were
reported to have greater abundance of CD14+ IL-+
and IFN-activated monocytes compared to healthy
controls [27]. These data suggest the potential risk of
inflammatory cytokine storms caused by monocytes.
In addition, ISGs upregulation in monocytes was
heterogeneous, with no clear relevance of COVID-19
severity [24].
2.2.1.2. Macrophages and DCs
Proinflammatory monocyte-derived macro-
phages had higher abundance in the BALF from
severe COVID-19 than mild cases [28]. Further
analysis of the heterogeneity of macrophages in
severe and mild patients found that the mild patients
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
389
highly expressed FABP4, while the severe patients
highly expressed FCN1 and SPP1. These data suggest
the imbalance of lung macrophage populations in
COVID-19 patients. In general, a highly
proinflammatory macrophage microenvironment is
present in the lungs of severe COVID-19 patients,
which may contribute to recruitment of other innate
immune cells and tissue damage. Additional data
showed that macrophages in the lower airways had a
stronger inflammatory signature than those within
the upper airways [29]. Conventional DCs (cDCs) and
plasmacytoid DCs (pDCs) have been reported to
significantly decrease in BALFs of patients with
severe COVID-19 [28]. Sánchez-Cerrillo et al. found
that CD1c+ cDCs preferentially migrated from blood
to lungs in severe COVID-19 cases, whereas CD141+
cDCs and CD123hi pDCs were depleted from blood
and were also absent in the lungs [30]. Another study
found a decrease in the resting DCs and an increase in
the activated DCs in the lungs of COVID-19 cases [24].
2.2.1.3. NK cells
Accumulating studies have reported low levels
of NK cells in the peripheral blood of COVID-19
patients with moderate and severe disease [24, 31, 32].
However, two reports assessing the immune cells in
BALF of COVID-19 patients have revealed that NK
cells are increased at this infection site [28, 29].
Maucourant et al. found low numbers but a strong
activation phenotype (Ki-67+, HLA-DR+, CD69+) of
NK cells in peripheral blood of COVID-19 patients
and adaptive NK cells, with high expression of
perforin, NKG2C, and Ksp37, was increased in
circulation of patients with severe COVID-19 [33].
Analysis of NK cell transcriptomic signatures
furthermore confirmed that the increased expression
of cytotoxic marker PRF1 and repair marker DDIT4 in
NK cells was associated with recovered COVID-19
patients [34]. In contrast, another study found
NKG2A, an inhibitory receptor, has been upregulated
on peripheral NK cells, meanwhile, the expressions of
the activation markers CD107a, IFN-γ, IL-2, and TNFα
were decreased. These results suggest the functional
exhaustion of peripheral NK cells in COVID-19
patients [31]. Moreover, in convalescent patients, the
frequency of NK cells was restored [31, 35].
2.2.1.4. Complement system
The complement system is a key component of
innate immune response to clear pathogens and
serves as a danger-sensing alarming system. Patients
with COVID-19 have been reported with the
activation of the complement system, and intense
complement activation was related to severe disease
[36]. COVID-19 patients showed the complement
activation in their lung, sera, and other organ tissue
[37, 38]. Mannose-associated serine protease 2
(MASP-2) is the critical component for mediating
activation of the lectin pathway. Nucleocapsid protein
of SARS-CoV-2 was found to interact with MASP-2
[38], which led to a strong complement activation and
further aggravated lung injury. Complement
activation products in COVID-19 patients included
the classical/lectin (C4d), alternative (C3bBbP) and
common pathway (C3bc, C5a, and sC5b-9), the lectin
pathway recognition molecule mannose-binding
lectin (MBL). Among which, sC5b-9 and C4d were
significantly higher in patients with respiratory
failure than without respiratory failure [37] (Fig. 2).
Similarly, complement 6 and complement factor B, as
the components of complement system, have been
reported to be elevated in serum of patients with
COVID-19, suggesting that the complement activation
is an early immune response mechanism [39].
2.2.1.5. Cytokine storm
It was observed that the cytokine storm was
associated with disease severity of COVID-19. A
clinical investigation of 41 confirmed cases infected
with SARS-CoV-2 in China revealed that the initial
plasma concentrations of IL-1B, IL-1RA, IL-7, IL-8,
IL-9, IL-10, basic FGF, G-CSF, GM-CSF, IFN-γ, IP10,
MCP1, MIP1A, MIP1B, PDGF, TNFα, and VEGF were
higher compared to healthy individuals [21]. Further
analysis of both ICU and non-ICU patients found that
ICU patients exhibited higher plasma concentrations
of IL-2, IL-7, IL-10, G-CSF, IP10, MCP1, MIP1A, and
TNFα than non-ICU patients [21]. Similarly,
additional studies quantified cytokines and
chemokines in serum samples of COVID-19 patients,
and the results showed remarkable increases in
circulating levels of IL-6, IL-1RA, CCL2, CCL8
CXCL2, CXCL8, CXCL9, and CXCL16 [19]. The
significant increased chemokines are likely
responsible for the recruitment of neutrophil (CXCL1,
CXCL2, CXCL6) and monocyte (CCL2 and CCL8) to
the lungs [19]. However, it is worth mentioning that
whole blood RNA levels of cytokines were not always
consistent with protein plasma levels. For example,
IL-6 had not detected an increase at the transcriptional
level in the peripheral blood of COVID-19 patients,
but there is a large amount of IL-6 protein. TNFα was
only moderately upregulated at the transcriptional
level, while circulating TNFα was significantly
increased [40, 41].
The interferon system plays an important role in
antiviral immunity, and it is critical to understand the
IFN response in COVID-19. An immune analysis
based on a cohort of 50 COVID-19 patients revealed
that type I IFN response was high in mild to moderate
patients, whereas severe patients had significantly
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
390
lower type I IFN responses than mild-to-moderate
patients [41]. In contrast to severe influenza,
COVID-19 patients exhibited a hyperinflammatory
profile in their peripheral blood mononuclear cells,
with a particular upregulation of TNF/IL--driven
inflammatory responses, and type I IFN responses
coexist with TNF/IL--driven inflammation in
monocytes from severe COVID-19 patients [42]. These
observations were consistent with an enhanced
expression of ISGs in patients with COVID-19 [41, 43].
Zhou et al. reported that the expression of 83 ISGs was
significantly upregulated in COVID-19 patients,
including IFITMs with direct antiviral activity,
compared with community-acquired pneumonia
patients [18]. However, in in vitro infection model
using cell lines, SARS-CoV-2 only induced low levels
of type I and type II IFNs, as well as moderate levels
of ISGs and a distinct proinflammatory cytokine
profiles including IL-1B, IL-6 and TNF, and many
chemokines (CCL20, CXCL1, CXCL2, CXCL3, CXCL5,
CXCL6 and CXCL16) [19, 40], resulting in efficiently
restricting SARS-CoV-2 infection. A longitudinal
immunological analysis of COVID-19 revealed that
IFNα and IFNλ in moderate patients were high levels
during the early phase and then declined, whereas
IFNα and IFNλ in severe patients showed a
continuous increase in overall [23]. The dynamic
changes of type I IFNs in COVID-19 patients suggest
that type I IFNs do not control the replication of
SARS-CoV-2 in vivo but are important drivers of the
disease progression.
2.2.2. Adaptive immune response
The adaptive immune response is critical for
controlling and eliminating most viral infections.
Studies of COVID-19 patients have observed that
adaptive immune response limits COVID-19 disease
severity and balanced CD4+ T cell, CD8+ T cell, and
antibody responses are protective, significantly
associated with milder disease. Understanding the
quantity and function changes in the three branches
(B cells, CD4+ T cells and CD8+ T cells) of adaptive
immune system in COVID-19 patients will provide
insights into immunity and pathogenesis of
SARS-CoV-2 infection, and the same knowledge also
contributes to the vaccine development and
evaluation of candidate vaccines.
2.2.2.1. Lymphopenia
Lymphopenia is prevalent in patients with
COVID-19 and is correlated with increased disease
severity. Patients who suffered COVID-19 showed
lower total blood lymphocyte compared to healthy
individuals, including a significant decrease in counts
of CD4+ T cells, CD8+ T cells, NK cells and NKT cells
[23]. The lymphocyte percentages were mostly lower
than 20% in severe patients, who were more likely to
exhibit lymphopenia than mild/moderate patients
[44]. Moreover, further analysis revealed that patients
with severe COVID-19 had a remarkable decrease in T
cells counts, but not B cells, especially CD8+ T cells
compared with moderate patients. These clinical data
suggest that lymphopenia can be used as one of the
effective indicators of disease severity and prognostic
evaluation in COVID-19 patients.
2.2.2.2. CD4+ T cells
Studies have noted that circulating
SARS-CoV-2-specific CD8+ and CD4+ T cells existed in
70% and 100% of convalescent COVID-19 patients
respectively [45], indicating that almost all
SARS-CoV-2 infections elicit the T cell response,
especially CD4+ T cell responses. Structural proteins
(spike, membrane, and nucleocapsid) of SARS-CoV-2
are the prominent targets of SARS-CoV-2-specific
CD4+ T cells. Additional CD4+ T cell responses also
against NSP3, NSP4, open reading frame (ORF) 3a,
and ORF8 [45]. SARS-CoV-2-specific CD4+ T cells can
be detected as early as 2-4 days after the onset of
symptoms [46, 47], which occurred in mild COVID-19
patients, accelerating viral clearance. Conversely, the
delayed appearance of SARS-CoV-2-specific CD4+ T
cells was associated with severe or fatal COVID-19 (>
22 days after the onset of symptoms in some cases)
[46, 47]. Moreover, SARS-CoV-2-specific CD4+ T cells
in severe COVID-19 patients displayed low antigen
avidity and clonality [48].
Virus-specific CD4+ T cells have the capacity for
differentiation into multiple helper and effector cell
types in response to SARS-CoV-2, which recruit
innate cells and provide help to B cells and CD8+ T
cells, with the abilities of direct antiviral activities and
facilitating tissue repair. Studies have been revealed
that the SARS-CoV-2-specific CD4+ T cells from both
acute and convalescent COVID-19 patients mainly
produced IFN-γ, TNF and IL-2, the classical cytokines
signature during type I T helper (Th1) cell responses,
with direct antiviral functions [45, 46, 49]. Another
branch of CD4+ T cells in immune responses against
SARS-CoV-2 is the differentiated T follicular helper
cells (Tfh), whose primary function is to assist B cells
in proliferation and neutralizing antibody production,
participating in humoral immunity. It has been found
that SARS-CoV-2-specific circulating Tfh cells (cTfh)
were a substantial fraction of the SARS-CoV-2-specific
CD4+ T cells in acute and convalescent COVID-19
patients [46]. And, the higher cell frequencies of
SARS-CoV-2-specific cTfh have been associated with
lower disease severity [46]. Cytotoxic T helper cells
(CD4-CTLs) are related cells with direct cytotoxic
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
391
activity. Besides cytotoxicity-associated transcripts,
the SARS-CoV-2-specific CD4-CTL were highly
enriched for transcripts encoding for a range of
chemokines such as CCL3, CCL4 and XCL1. These
chemokines participate in the recruitment of myeloid
cells (neutrophils, monocytes, and macrophages), NK
cells, and DCs to the sites of viral infection [50].
According to the relevant data, the proportion of
CD4-CTLs and antigen-specific regulatory T cells
(Tregs) exhibited an obvious negative correlation [51].
2.2.2.3. CD8+ T cells
CD8+ T cells are important in clearing viral
infections due to their capacity for killing infected
cells. In COVID-19 patients, the presence of
SARS-CoV-2-specific CD8+ T cells has been related to
less severe disease [46]. Similarly to SARS-CoV-
2-specific CD4+ T cells, SARS-CoV-2 CD8+ T cells are
specific for multiple SARS-CoV-2 antigens, such as
spike, nucleocapsid, membrane and ORF3a protein
[45]. SARS-CoV-2-specific CD8+ T cells can be
detected as early as day 1 post-symptom onset [52].
The data showed that SARS-CoV-2-specific CD8+ T
cells were detected in 87% of convalescent COVID-19
cases but only in 53% of acute cases [46]. In the acute
COVID-19 cases, SARS-CoV-2-specific CD8+ T cells
were characterized by activation marker (CD38,
HLA-DR, and Ki-67) and predominantly expressed
IFN-γ, granzyme B, perforin, and CD107a, with
cytotoxic effector functions [45, 53]. However,
SARS-CoV-2-specific CD8+ T cells in convalescent
COVID-19 patients tended to an early differentiated
memory phenotype (CCR7+ CD127+ CD45RA-/+
TCF1+) [54]. From the foregoing, increasing the
number of CD4 and CD8 T cells through the
combination of other therapeutic approaches may be a
better method for the treatment of COVID-19 patients,
including the use of drug (such as CD3 antibody and
CD28 antibody [55]) and drug-free (such as
phototherapy [56]) therapeutic strategies.
2.2.2.4. Antibody response
The SARS-CoV-2 infected individuals exhibited
seroconversion within 20 days after symptom onset
[24, 57, 58]. Seroconversion for immunoglobulin (Ig)
M and IgG antibodies against SARS-CoV-2 occurred
simultaneously or sequentially [57]. A study based on
173 COVID-19 patients revealed that the
seroconversion rates for total antibodies, IgM, and
IgG were 93.1%, 82.7%, and 64.7%, respectively [59].
The spike and nucleocapsid of SARS-CoV-2 are
primary antigens testing for seroconversion [60].
Spike is the main target of neutralizing antibodies
against SARS-CoV-2, and its receptor binding domain
(RBD) is the target of more than 90% of neutralizing
antibodies in COVID-19 patients [60, 61], whereas
some other neutralizing antibodies instead target the
NTD [61]. Moreover, most COVID-19 patients had
detectable IgG and IgA responses to spike and
nucleocapsid, whereas IgM responses were limited to
spike and undetectable for nucleocapsid [60].
Among most SARS-CoV-2-infected patients,
neutralizing antibodies developed rapidly. Zhao’s
data showed that the antibodies were less than 40%
within 1 week since onset, and could rapidly increase
to 100.0% (total antibodies), 94.3% (IgM), and 79.8%
(IgG) within 15 days [59]. And neutralizing antibodies
titers stayed relatively stable for at least 5 months [62].
Additionally, it is worth noting that SARS-CoV-2
neutralizing antibody titers are positively correlated
with COVID-19 disease severity [57, 60, 63].
Consistently, most convalescent cases who recover
from COVID-19 do not exhibit high levels of
neutralizing antibodies activity. These antibodies had
little somatic hypermutation and were highly
enriched for the usage of gene VH1-69, VH3-30-3, and
VH1-24 [61]. Furthermore, the production of
SARS-CoV-2 neutralizing antibodies did not require
affinity maturation. Collectively, these data suggest
that the neutralizing antibodies against SARS-CoV-2
are relatively easy to generate.
2.2.2.5. Immune memory
Immunological memory is composed of memory
CD4+ T cells, memory CD8+ T cells and memory B
cells. All three major types of immune memory are
substantially generated after SARS-CoV-2 infection.
About 95% of individuals kept immune memory at ~6
months after SARS-CoV-2 infection [64].
About 90% of subjects are seropositive for
SARS-CoV-2 neutralizing antibodies at 6 to 8 months
after symptom onset [64]. SARS-CoV-2 RBD IgM and
IgG titers steadily decreased within 1 to 6 months
after infection, whereas IgA was less declined
relatively [65]. One cross-sectional analysis of
COVID-19 subjects has revealed that SARS-CoV-2
specific memory B cells were detectable in all subjects
at 6 months post-infection. Moreover, the frequencies
of memory B cells specific to spike, RBD, and
nucleocapsid increased over time in the following 4 to
5 months post-infection and then plateaued [64]. RBD
memory B cells had undergone affinity maturation
after infection 6 months and expressed the increased
potency neutralizing antibodies, indicating
continuous evolution of humoral immunity [65].
The memory CD4+ T cells predominantly
consists of Th1 and Tfh cells [64, 66]. The majority of
SARS-CoV-2 memory CD8+ T cells were mainly
terminally differentiated effector memory cells
(TEMRA), with small populations of central memory
(TCM) and effector memory (TEM) [64]. A few studies
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
392
have assessed memory CD4+ T cells and CD8+ T cells
at least 6 months after infection. One study found that
~90% of subjects were positive for SARS-CoV-2
memory CD4+ T cells and ~50% were positive for
memory CD8+ T cells after 6 months primary infection
[64]. There was a steady decline in SARS-CoV-2
memory CD8+ T cells and CD4+ T cells within 8
months after infection.
2.3. COVID-19 on blood circulation system
2.3.1. Hematologic Parameters of COVID -19
Patients with COVID-19 usually present with
abnormal hematologic changes, including white
blood cells reduction, lymphopenia, and
thrombocytopenia. Based on the clinical data of 1099
COVID-19 cases provided by Guan et al. [67], the vast
majority of patients on admission presented with
lymphocytopenia (83.2%), whereas 36.2% had
thrombocytopenia, and 33.7% showed leukopenia.
And these hematological abnormalities were more
prominent in severe versus non-severe cases.
Several factors may contribute to COVID-19
associated lymphopenia. On the one hand, previous
study showed that lymphocytes expressed the ACE2
receptor on their surface [6], thus SARS-CoV-2
directly infected those cells and eventually led to cell
lysis. On the other hand, the virus particles
disseminate through the respiratory mucosa and
infect other cells, inducing a cytokine storm in the
body. It is characterized by markedly increased levels
of interleukins (IL-2, IL-6, IL-8), G-CSF, IFN-γ
inducible protein (IP-10) and TNFα, which promote
lymphocyte apoptosis [68-70]. Furthermore, massive
cytokine activation may be also associated with
atrophy of lymphoid organs, including the spleen,
and further impairs lymphocyte turnover [71].
2.3.2. COVID-19 and cardiovascular system
Patients with prior cardiovascular disease are at
higher risk for adverse events from COVID-19.
Individuals without a history of cardiovascular
disease are at risk for incident cardiovascular
complications [72]. The vast majority of COVID-19
patients with previous cardiovascular disease
presented with cardiovascular complications
including hypertension (40%) [73, 74], whereas 10%
had coronary heart disease, 17% had cardiac
arrhythmias, and 4% showed heart failure [22, 75].
Patients with more severe clinical presentations
present with comorbidities such as hypertension
(58%), heart disease (25%), and arrhythmias (44%) [76,
77]. Additionally, cardiovascular manifestations,
during COVID-19, are mostly represented by acute
cardiac injury (ACI), defined as a rise of cardiac
troponin values, with or without ejection fraction
decline or electrocardiographic abnormalities. The
prevalence of acute cardiac injury among COVID-19
patients was 10-23% [21].
Furthermore, COVID-19 patients are also at an
increased risk of venous thromboembolism and their
main coagulation parameters (elevated D-Dimer
levels, fibrin degradation products) are also altered,
especially in patients with severe manifestations [72].
The direct effects of COVID-19 or the indirect effects
of infection, such as severe illness and hypoxia, may
predispose patients to thrombotic events. Preliminary
reports suggest that hemostatic abnormalities,
including disseminated intravascular coagulation
(DIC), occur in patients affected by SARS-CoV-2 [78,
79]. Additionally, the severe inflammatory response,
critical illness, and underlying traditional risk factors
may all predispose to thrombotic events [79, 80].
Overall, patients with cardiovascular disease
represent more than 20% of all fatal cases, with a case
fatality rate of 10.5% [75].
2.4. COVID-19 on nervous system
Similar to other coronaviruses (mainly
SARS-CoV-1, MERS-CoV and OC43), the possibility
of SARS-CoV-2 invading the central nervous system
has been proposed [81]. In vivo studies in human
ACE2 transgenic mice have shown that SARS-CoV-2
can infect neurons and cause neuronal death in an
ACE2 dependent manner [82]. In brain cells derived
from human pluripotent stem cells, dopaminergic
neurons (rather than cortical neurons or microglia) are
particularly sensitive to SARS-CoV-2 infection [83].
However, the results of clinicopathological studies on
the detection of virus in brain or cerebrospinal fluid
(CSF) are different. Some studies have shown that
SARS-CoV-2 RNA exists in the brain or cerebrospinal
fluid of patients with encephalopathy or encephalitis
after death, but the level is very low [84]. Other
studies have failed to detect viral invasion, even if
there is evidence of CSF inflammation [85, 86]. Recent
studies have shown that SARS-CoV-2 can enter the
brain through damaged blood-brain barrier (BBB),
olfactory bulb/olfactory nerve or lymphatic vessels
[87]. SARS-CoV-2 can infect and directly cause
endothelial cell damage, increase BBB permeability,
and cause edema formation. It has been reported that
ACE2 is expressed in physiological conditions in
neurons and glial cells, making the brain more
vulnerable to SARS-CoV-2 infection [88]. Infected
neurons release inflammatory molecules and activate
nearby immune cells, including mast cells, endothelial
cells, pericytes, neurons, microglia and astrocytes.
When ACE2 is expressed in brain endothelial cells,
SARS-CoV-2 infection can cause cerebral hemorrhage
and BBB dysfunction [89], leading to endothelial cell
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
393
damage, brain edema formation, neuronal death and
cognitive decline (Fig. 3). In addition, researches
demonstrated that ACE2 is also expressed in
pericytes, another apart of mural cells of the central
nervous system (CNS). SARS-CoV-2 may directly
destroy pericyte and further reduce blood supply to
the brain, and ultimately result in neuronal
dysfunction [90, 91]. Pericytes are perivascular cells
within the brain that are proposed as SARS-CoV-2
infection points [92]. Moreover, the infection of
olfactory bulb pericytes may disrupt neuronal signals
through local inflammation and cytokine release,
which are related to the functional impairment caused
by olfactory bulb vascular injury and hypoperfusion
[92]. A recent study shows that SARS-CoV-2 could
preferentially infect astrocyte over other cells [93].
They exposed the brain organoids to SARS-CoV-2 and
found that almost only astrocytes were infected.
Similarly, Daniel Martins-de-Souza et al. analyzed the
brains of 26 COVID-19 deceased, among 5 samples in
which the brain cells were infected with the
SARS-CoV-2, up to 66% of the infected cells were
indeed astrocytes [94].
Severe neurological complications in COVID-19
are both rare and variable in nature. Indeed, any part
of neuraxis seems to be susceptible to damage by
SARS-CoV-2. Neurological disorders may result from
systemic cardiopulmonary failure and metabolic
abnormalities caused by infections, direct viral
invasion, or viral autoimmune reactions. In an
unpublished report from Wuhan, China, 78 (36.4%) of
214 hospitalized COVID-19 patients had some form of
neurological disorder, the most common features of
which were dizziness, headache, hypogeusia, and
hyposmia [95]. While serious neurological
complications have been reported in patients with
otherwise mild COVID-19 [96], the most severe
complications occur in critically ill patients and are
Figure 3. Brain entry of SARS-CoV-2. Circulating SARS-CoV-2 and cytokines act on endothelial cells, leading to inflammation and BBB opening. Once in the perivascular space,
these factors induce inflammation in vascular parietal cells, microglia and macrophages resident in the brain. The cytokines may affect the function of neurons and lead to cytokine
sickness, which is a potential cause of COVID-19 encephalopathy.
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
394
associated with significantly higher mortality [83, 97].
And neurological disorders were found to be more
common in critically ill patients, such as stroke in 6
(2.8%), impaired consciousness in 16 (7.5%), and
muscle injury in 23 (10.7%). These were subdivided
into those thought to reflect CNS and peripheral
nervous system (PNS) [95].
2.4.1. Central nervous system dysfunction
2.4.1.1. Dizziness and headache
Headache is a possible symptom in any systemic
viral infection such as SARS-CoV-2. In COVID-19,
headache usually coincides with fever. Headache was
reported in 6.5-14.1% of COVID-19 cases [73, 98, 99].
However, the prevalence of headaches in COVID-19
infection seems to be underestimated in terms of
variety and clinical description. According to this
finding raised by Dr. Robert Belvis [100], headaches
related to COVID-19 can be classified in the 2 phases
of the disease. Acute headache, primary cough
headache, tension type headache, and heterologous
headache caused by systemic viral infection can
appear in Phase I (influenza like phase) and hypoxia
and headache caused by new onset headache can
occur in Phase II (cytokine storm phase with an
increase of IL-2, IL-6, IL-7, IL-10, TNFα, G-CSF, IFN-γ
inducible protein 10, MCP1, and macrophage
inflammatory protein 1-α). Headache may be
associated with cytokine storm in SARS-CoV-2
infection, but further research is needed to better
understand this link.
2.4.1.2. Acute cerebrovascular disease
With the spread of COVID-19 around the world,
there is more and more evidence related to
cerebrovascular diseases and other forms of vascular
diseases. COVID-19 cerebrovascular disease appears
to be predominantly ischemic and involves large
vessels. In the elderly, it reflects the underlying
severity of the systemic disease as well as the
hyperinflammatory state, whereas in the younger
patients, it seems to be due to hypercoagulopathy
[101, 102]. Besides hypercoagulable state,
SARS-CoV-2 can also infect and damage endothelial
cells. However, it remains to be determined whether
the endothelial cell damage caused by the virus or
even the real vasculitis will lead to the
cerebrovascular syndrome associated with COVID-19,
which will require more detailed angiography and
neuropathological analysis.
Stroke may have significant interaction with
COVID-19, and stroke is not uncommon among
patients hospitalized with COVID-19. Several studies
have reported strokes in COVID-19 patients, with
rates ranging from 1%-3% and up to 6% of critically ill
patients [5, 73, 99, 103]. These patients may develop
more severe coagulopathy, defined as
COVID-19-related coagulopathy, which may be
caused by inflammation, including inflammatory
cytokine storms. It is not clear whether these strokes
are caused by SARS-CoV-2 or the incidence rate of
stroke in these high-risk groups, and these high-risk
groups also happened to have SARS-CoV-2.
SARS-CoV-2 infection does play a role in stroke,
which is reasonable, because infection usually
increases the risk of stroke. Therefore, understanding
the relationship between infection and stroke has
taken on urgency in the era of the COVID-19
pandemic. The association of infection and stroke is
also bidirectional. Furthermore, ACE is known for its
role in blood pressure regulation through the renin
angiotensin aldosterone system (RAAS), and it can
also play a role in fertility, immunity, hematopoiesis
and obesity, fibrosis and Alzheimer’s dementia [104].
More importantly, it is a functional receptor of
SARS-CoV-2 [105]. Therefore, understanding the
interaction between SARS-CoV-2 and ACE2 is very
important for the design of therapy for this disease. In
the case of severe infection, myocardial injury and
arrhythmias, such as atrial fibrillation, can lead to
cardiac embolism and cerebral infarction [106]. In
addition to primary viral diseases, a considerable
number of critical patients with COVID-19 may also
have secondary bacteremia. In one case series, about
10% of patients requiring mechanical ventilation have
bacteremia, which increases the risk of stroke by more
than 20 fold [107]. Septic cerebral emboli often lead to
hemorrhage. In a postmortem magnetic resonance
imaging (MRI) study, 10% of brain showed signs of
hemorrhage [108]. Emerging evidence suggested the
role of infection, as a contributor to long-term risk of
atherosclerotic disease and stroke; immune
dysregulation after stroke and its effect on the risk of
stroke-associated infection; and the impact of
infection at the time of stroke on the immune reaction
to brain injury and subsequent cognitive decline [109].
In conclusion, these clinical findings suggest that
SARS-CoV-2 may have adverse effects on brain
through a variety of pathophysiological pathways
and eventually lead to vascular brain injury.
2.4.2. Peripheral nervous system
2.4.2.1. Peripheral organ dysfunctions
COVID-19 also damages other organs, and
metabolic and pathological evidences on
COVID-19-induced renal, cardiac, hepatic,
gastrointestinal and endocrine organ damage have
been presented so far [106, 107, 110, 111]. The
resulting systemic metabolic changes, including water
and electrolyte imbalance, hormonal dysfunction, and
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
395
accumulation of toxic metabolites, may also
contribute to some of the nonspecific neurological
manifestations of the disease, like confusion,
agitation, headache, cardiac involvement, which may
affect brain by reducing cerebral perfusion. The lung
is the most seriously affected organ in COVID-19,
accompanied by a large number of alveolar injuries,
edema, inflammatory cell infiltration, microvascular
thrombosis, microvascular injury and bleeding [112].
SARS-CoV-2 was mainly detected in lung cells and
epithelial progenitor cells [112, 113]. Severe hypoxia
(acute respiratory distress syndrome, ARDS) caused
by respiratory failure caused by lung injury requires
auxiliary ventilation [114]. Consistent with hypoxic
brain damage, the autopsy study of COVID-19
showed that neuronal damage was found in the most
vulnerable areas of brain, including neocortex,
hippocampus and cerebellum [84, 115].
2.4.2.2. Neurological autoimmune disorders with
COVID-19 on PNS
Viral illnesses may trigger an autoimmune
response, which affects the central or peripheral
nervous system. As mentioned earlier, acute
inflammatory demyelinating peripheral neuropathy
(AIDP)/Guillain-Barre syndrome (GBS) may be a
consequence of infection of peripheral nervous
system. MERS also causes post infectious brainstem
encephalitis and GBS [116]. Reports of SARS-CoV-2
transverse myelitis also began to appear [73]. GBS is
an acute polyradiculopathy characterized by rapid
progressive symmetrical limb weakness, sensory
disturbance during examination, and facial weakness
in some patients. 11 patients had GBS with weakness
of all four limbs with or without sensory loss
[117-119], of which three patients only had a paralytic
variant with leg weakness [118, 120, 121], and one had
lower limb paresthesia [118].
2.5. COVID-19 on motor system
Motor complications with COVID-19 such as
critical illness myopathy, polyneuropathy, GBS, Bell’s
palsy, and Parkinson’s disease (PD) have been
reported recently [122].
2.5.1. Autonomic dysfunction preceding acute motor
axonal neuropathy (AMAN)
A 20-year-old patient had previous autonomic
dysfunction characterized by sinus arrhythmia,
postural hypotension, intermittent sweating,
constipation, erectile dysfunction, and chest crush
[123]. This is unique in this case, except for the fact
that autonomic dysfunction precedes motor
weakness, which is very rare. Since the disease
associated with COVID-19 was mild and almost
asymptomatic, the patient was only treated with
acetaminophen. For AMAN, intravenous
immunoglobulin (0.4 g/kg/day, for 5 consecutive
days) was started. After 15 days of treatment, the
patient's dyskinesia and autonomic nervous system
began to improve. After rigorous physical therapy,
the patient was able to walk with some help one
month after admission.
2.5.2. Oculomotor nerve palsy and motor peripheral
neuropathy
A report described a 65 year old man with
oculomotor nerve palsy [124]. He had a 5-day history
of persistent diplopia and left blepharoptosis. MRI
and Magnetic resonance angiography (MRA) showed
no abnormalities, but computed tomography (CT) of
chest showed diffuse ground glass opacity.
SARS-CoV-2 was detected in throat swabs.
Importantly, a case reported that a 69 year old man
was admitted to the COVID-19 ward with bilateral
weakness of lower limbs three days before admission
[120]. Due to chronic cough, he underwent a
COVID-19 swab test in the emergency room and was
admitted to the COVID-19 ward. The strength of his
bilateral knee extension was reduced by four fifths,
the strength of other muscle groups was normal, his
knees and ankles had no convulsions, and his gait was
ataxia. This case is different from the case of GBS
reported in the Journal of Neurology of the Lancet.
There is microbiological evidence of COVID-19
infection at the time of admission, and influenza like
symptoms appear only 7 days after the symptoms
appear. He had distal weakness and hyporeflexia, no
back pain or sensory level, suggesting motor
peripheral neuropathy. However, this diagnosis still
needs further characterization and analysis. At
present, the physical condition of the case does not
allow us to carry out further characterization.
2.5.3. Bell’s palsy
Wan et al. described the first case of Bell’s palsy
in a 65 year old woman who developed left lower
motor neuron facial paralysis 2 days after mastoid
pain. Interestingly, the patient had no symptoms of
other viral diseases [125]. MRI showed no
abnormality, but CT showed a piece of ground glass
shadow in the right lower lung, which was suspected
to be SARS-CoV-2. SARS-CoV-2 infection was
confirmed by real-time reverse transcription
polymerase chain reaction (RT-PCR).
2.5.4. Parkinson’s disease and motor symptom
Patients with underlying neurological
dysfunction, such as PD, often have associated
cardiovascular and respiratory disorders, which
increase their risk of developing severe COVID-19. In
addition, due to fever and altered dopaminergic drug
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
396
intake, PD patients with COVID-19 have developed
Parkinson’s disease high fever syndrome, a motor
disorder emergency [126]. Although patients who
experience this phenomenon may recover from
COVID-19, some patients may leave significant
disabilities, while others may not survive [127].
Recently, people with or without PD participating in
the online study Fox Insight (FI) were invited to
complete a survey to assess COVID-19 symptoms and
the pandemic’s effect. The report showed that people
with PD and COVID-19 experienced new or
worsening motor (63%) symptom. And people with
PD but not diagnosed with COVID-19 reported
disrupted medical care (64%), exercise (21%), and
social activities (57%), and worsened motor (43%)
symptoms [128]. These results suggested that during
COVID-19 infection, people with PD reported
worsening of many PD-related motor and non-motor
symptoms, including rigidity, tremor, walking
difficulties, emotional symptoms, cognition and
fatigue.
2.6. COVID-19 on urinary system
While most COVID-19 patients present with
mild symptoms, a small percentage can gradually
develop acute respiratory distress syndrome and
multiple organ dysfunction syndromes, leading to
death [129]. Because these symptoms may overlap
with other common disease processes, it is difficult to
identify these symptoms as the underlying cause
directly related to COVID-19. Recently, Mumm and
his colleagues reported increased frequency of
urination, and identified this in seven males out of 57
patients currently being treated in COVID-19 wards
[130]. In the absence of any other cause, urinary
frequency may be secondary to viral cystitis due to
the underlying COVID-19 disorder. We recommend
considering urinary frequency as a memory tool for
patients with infectious symptoms to increase
urologists’ awareness during the current COVID-19
pandemic, preventing the fatal effects of erroneous
interpretation of urinary symptoms.
2.6.1. Lower urinary tract symptoms
One of the most frequently reported
epidemiological data is gender related COVID-19
mortality. Studies conducted in various countries
have shown that men are more susceptible to
COVID-19 infection. Male patients accounted for 73%
of deaths in China, 59% of deaths in Korea, and 70% of
patients who died in Italy were male [131, 132]. In
addition, a recent review of current epidemiological
studies of 59254 patients from 11 different countries
suggested an association between male sex and higher
mortality [133]. In addition, experimental studies
performed by Channappanavar and his colleagues
showed that male rats were more susceptible to
SARS-CoV infection than age-matched female rats
[134]. Epidemiological studies of COVID-19 are
essential to better understand the pathogenic
mechanisms of the disease and to identify good
treatment strategies. Lower urinary tract symptoms
(LUTS) are a term that encompasses a wide range of
symptoms that occur after storage and voiding. LUTS
is common in adult men and is often associated with
benign prostatic hyperplasia (BPH) [135]. The
prevalence of BPH increases significantly with age,
and BPH-related LUTS often emerged as a natural
result of aging and androgen exposure. Prostatic
hyperplasia increases urethral resistance, leading to
compensatory changes in bladder function. Because
bladder outlet resistance increases, detrusor pressure
also increases with increasing urine flow. Increased
detrusor resistance also causes LUTS by affecting
bladder storage function [136]. LUTS was assessed by
approved questionnaires, such as the International
Prostate Symptom Score (I-PSS), and the results
indicated that LUTS could effectively predict the
severity of COVID-19.
2.6.2. Acute kidney injury
Acute kidney injury (AKI) has been reported in
up to 25% of critically ill patients with SARS-CoV-2
infection, in particular in those with serious infections,
and has been associated with substantial morbidity
and mortality [137, 138]. In most studies, AKI
develops throughout hospitalization, with a mean of 5
to 9 days after admission [139, 140]. AKI develops
more frequently in patients with the most severe
diseases (especially ARDS, requiring invasive
mechanical ventilation), including elderly patients or
those with hypertension or diabetes. The causes of
kidney involvement in COVID-19 may be
multifactorial, and cardiovascular comorbidity and
predisposing factors (such as sepsis and nephrotoxin)
are important factors. However, renal tubular injury is
common and associated with the cytopathic effect of
renal resident cells and cytokine storm syndrome
[141, 142].
2.7. COVID-19 on reproductive system
Since the emergence of the SARS-CoV-2 infection
in December 2019, it has rapidly spread across all over
the world. Additionally, it has been demonstrated
that SARS-CoV-2 infection not only damage to
respiratory system, but other organs of human, such
as heart, liver, oesophagus, kidney, bladder, and
ileum [143]. As mentioned above, ACE2, the
functional receptor for SARS-CoV-2, modulates the
cleavage of Ang II and Ang 1-7. Because SARS-CoV-2
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
397
enters cells by binding to ACE2 receptor, the
reproductive cells and/or tissues expressing ACE2
may be susceptible to virus infection, and their
functions may be interfered theoretically. ACE2, Ang
II and Ang 1-7 can regulate the basic functions of male
and female reproductive system. In the female, it
includes folliculogenesis, steroidogenesis, oocyte
maturation, ovulation and endometrial regeneration
[144, 145]. In the male, testicular ACE2 may regulate
testicular function, play a role in sperm function, and
may affect sperm contribution to embryo quality [146,
147]. An important and interesting topic in the era of
COVID-19 is the ability of virus to affect male and
female reproductive capacity (Fig. 4), and whether
pregnant women with COVID-19 have an increased
risk of death or comorbidity.
2.7.1. Reproductive hormones
One of the main functions of ovary and testis is
steroid production. Therefore, the evaluation of sex
hormone levels can provide the evaluation of gonadal
function in patients with COVID-19. Ma et al.
compared sex related hormone levels in 119
reproductive men infected with SARS-CoV-2 with 273
age-matched controls [148]. Most patients have
moderate to severe diseases. The serum luteinizing
hormone (LH) was increased and serum Testosterone
(T)/LH ratio was decreased in the COVID-19 group.
Rastrelli et al. found that the deterioration of clinical
condition is accompanied by the gradual decrease of T
level and the increase of LH level [149]. However,
these results should be interpreted with caution, as
pre-infection sex hormone baselines are not available
for these patients. In addition, hypogonadism is a
common systemic disease. In the case of COVID-19, it
is not clear whether the low T level observed is the
result of the direct effect of COVID-19 on gonadal
function [150]. In the female, severe acute illness may
alter hypothalamic pituitary gonadal (HPG) axis
function, reducing endogenous estrogen and
progesterone production [151].
2.7.2. Sex and COVID-19
ACE2 receptor is more abundant in male
reproductive system than in female reproductive
system. ACE2 was low expressed in oviduct (ciliated
cells and endothelial cells), ovary, vagina, cervix and
endometrium [152, 153]. On the other hand, the
expression of ACE2 was the highest in testis, high in
leydig cells and sertoli cells, and medium in seminal
vesicle cells [154, 155]. Therefore, it is expected that
testis is more vulnerable to SARS-CoV-2 infection
than ovary.
2.7.2.1. Male
As we all know, ACE2 not only expresses in the
lung, but also extensively expresses in spermatogonia,
sertoli and leydig cells in testicle. Indeed, it has been
found that the testis may be infected with
SARS-CoV-2, which may lead to further reproductive
system diseases [156, 157]. Similarly, it is well known
that viruses can enter testicular cells, cause viral
orchitis, and even lead to male infertility and
testicular tumors [158, 159]. In ACE2 positive cells, the
abundance of transcripts related to SARS-CoV-2
replication and transmission was higher than that
related to male gametogenesis. The expression of
ACE2 in human testis suggests that SARS-CoV-2 may
infect male gonads and lead to male reproductive
dysfunction. We know that temperature is very
important for cell growth and development, and 37°C
is the best temperature for cell growth. However,
almost all SARS-CoV-2 infected people have
persistent fever. When the human body is in a state of
Figure 4. Effects of COVID-19 on the male and female reproductive systems.
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
398
high fever for a long time, the testicular temperature
will change, and the germ cells will be damaged and
degenerated [160, 161]. Furthermore, several results
showed that SARS-CoV-2 existed in semen and testis
of SARS-CoV-2 infected patients in acute and
convalescent stages. Therefore, it also supports the
result that SARS-CoV-2 can be sexually transmitted
through men [162, 163]. SARS-CoV-2 infection can
cause systemic local inflammation, and due to the
imperfect blood testis/vas deferens/epididymis
barrier, SARS-CoV-2 may propagate to the male
reproductive tract [164]. However, the virus cannot
replicate in the male reproductive system, it may
persist, and this phenomenon may be due to the
special immunity of testis [165, 166]. These results
suggest that sexual transmission may be an important
link in the prevention of transmission. A recent report
on 31 Italian male COVID-19 patients noted that some
patients developed hypergonadotropic hypogona-
dism after disease onset [149]. In this study, low levels
of serum testosterone (total and free) may serve as a
predictor of poor outcome in SARS-CoV-2-infected
men. Testosterone, as a regulator of endothelial
function, inhibits inflammatory responses by
increasing the levels of anti-inflammatory cytokines,
such as IL-10, and decreasing the levels of
pro-inflammatory cytokines, such as TNFα, IL-6, and
IL- [167, 168]. Therefore, it can be hypothesized that
suppressed testosterone levels may be one of the
reasons for the large differences in mortality and
hospitalization between men and women and may
also explain why SARS-CoV-2 most commonly infects
elderly men. As mentioned earlier, patients with
hypogonadism have higher concentrations of TNFα,
IL-6 and IL- due to reduced inhibition. This
eventually worsens endothelial dysfunction and
further impairs erectile function. While erection is
certainly a trivial matter for patients in ICU, there is
reason to suspect that impaired vascular function may
persist among survivors of COVID-19 and even
become a public health problem in the coming
months. In addition, since erectile function is a
predictor of heart disease [169, 170], investigating
whether erectile dysfunction occurs in patients with
COVID-19 may also be a good alternative indicator of
general cardiovascular function, improving patient
care and quality of life. In addition, only a small
retrospective study assessed the presence of
SARS-CoV-2 in prostate secretions. One study
evaluated prostate secretions from 18 male patients
with confirmed COVID-19 and 5 suspected cases, and
found that no SARS-CoV-2 RNA expression was
detected in the samples of all patients assessed [171].
As any kind of sudden disease, there are more doubts
and hypotheses, rather than certainty, about the
impact of COVID-19 on male reproductive system.
Many studies have been carried out to better
understand the disease and its short- and long-term
effects on health. As has been demonstrated in other
viral diseases, the involvement of male reproductive
system is a possibility, which may reveal a new route
of transmission and/or impact on its function. The
virus has been found in the semen of infected patients,
but its impact on male reproductive health remains to
be further investigated.
2.7.2.2. Female
Studies have shown that ACE2 mRNA is highly
expressed in the ovaries of women of childbearing age
and postmenopausal women. These results implied
that female reproductive system may be at risk of
SARS-CoV-2 infection [172]. Single cell sequencing
was used to analyze the expression of ACE2,
TMPRSS2, cathepsin B and L (CTSB and CTSL,
respectively) in human ovarian cells [173]. It was
found that the expressions of ACE2 in stromal cells
and perivascular cells of ovarian cortex were very
low. TMPRSS2 was not expressed in different types of
oocyte nest cells, while CTSB and CTSL were
expressed in all ovarian cells. There was no
co-expression of ACE2/CTSB and ACE2/CTSL in all
ovarian cell types. Because ACE2 needs the
co-expression of protease TMPRSS2 or CTSB/L to
make protein on its surface and to ensure enter host
cells. The expression rate of ACE2 in ciliated cells,
secretory cells and leukocytes was less than 5%. On
the contrary, the expression levels of protease
TMPRSS2 and CTSL/B were different in different
fallopian tube cells, but CTSL was not detected in any
fallopian tube cells. No co-expression of ACE2,
TMPRSS2 or CTSB was observed in any oviduct cells.
Furthermore, studies have shown that 70
pregnant women infected with SARS-CoV-2 had
symptoms such as fever (84%), cough (28%), and
dyspnea (18%). Obstetric complications included
preterm delivery (39%), intrauterine growth
restriction miscarriage (10%) and miscarriage (2%).
Amniotic fluid, cord blood, throat swabs and neonatal
milk were collected from 9 pregnant women with
SARS-CoV-2 infection in Wuhan, China, and the
results indicated that there was no direct evidence of
vertical transmission of SARS-CoV-2 [174, 175]. But it
was found in additional studies that ACE2 has
transient overexpression and increased activity
during pregnancy, particularly in the placenta,
implying that there may be vertical transmission.
Previous clinical studies have not observed evidence
of vertical transmission of SARS-CoV-2 among cases,
a phenomenon that still needs to be more carefully
investigated in clinical practice [176].
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
399
SARS-CoV-2 may infect ovary, uterus, vagina
and placenta through the universal expression of
ACE2. In addition, SARS-CoV-2/ACE2 may interfere
with female reproductive function, leading to
infertility, menstrual disorders and fetal distress [177].
We recommend following up and evaluating fertility
after recovery from SARS-CoV-2 infection and, if
possible, postponing pregnancy, especially in young
women. Moreover, we should continue to pay
attention to the situation of pregnant patients and
fetus, and take timely measures. In order to reduce the
incidence of SARS-CoV-2 infection, special care was
given to healthy pregnant women, puerpera and
newborns.
COVID-19 sex differences in incidence,
comorbidities, and mortality males are at higher risk
and require prompt action to understand the sources
of biological and behavioral differences. As the impact
of SARS-CoV-2 on male/female reproductive system
becomes more intensively studied, we will control
and prevent the SARS-CoV-2 infection system in the
male/female reproductive system more and more
effectively.
2.8. COVID-19 on digestive system
At the initial stage of COVID-19 pandemic, a
series of respiratory manifestations caused by the
virus are the first to be discovered and concerned.
Thence, SARS-CoV-2 is initially considered to be most
likely to cause respiratory illnesses and spread from
human to human mainly via respiratory tract. Since
SARS-CoV-2 RNA in stool specimen was first
reported in the study of the first case of COVID-19
infection in USA [178]. Ongoing reports of viral
gastrointestinal infection and fecal-oral transmission
of the virus are a matter of widespread concern.
2.8.1. The impact of SARS-CoV-2 on gastrointestinal
tract
The digestive system has been reported not only
as a site of disease expression, but also as a possible
driver of disease severity and viral transmission.
Several studies have reported a high prevalence of
gastrointestinal symptoms in patients infected with
SARS-CoV-2, mainly including loss of appetite,
nausea, vomiting, diarrhea and abdominal pain
[178-180] (Fig. 5). Importantly, multiple studies have
confirmed a fraction of COVID-19 patients only
experience abdominal symptoms without fever or
respiratory manifestations [179, 180]. Therefore, it is
important that clinicians should maintain a high index
of vigilance in patients with gastrointestinal
symptoms. In addition, SARS-CoV-2 appears to
persist in patients’ stool specimen, even though the
patients’ throat swabs become negative [181, 182]. The
study also concluded that patients with
gastrointestinal symptoms have significantly longer
time from onset to hospital admission than patients
without gastrointestinal symptoms [180]. In contrast,
a study reported a low incidence (3.8%) of
gastrointestinal symptoms [67]. Of course, due to the
difference in medical record samples and symptom
ascertainment, different conclusions are normal.
However, the risk of virus transmission here is
unanimously recognized and cannot be ignored.
Figure 5. SARS-CoV-2 infection of the gastrointestinal tract.
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
400
Endoscopic and histologic findings give us a
more comprehensive understanding of the impact of
SARS-CoV-2 on the gastrointestinal tract. In the study,
a patient with COVID-19 present upper gastro-
intestinal bleed [181]. Gastrointestinal endoscopy was
performed and mucosa damage in the esophagus was
observed [181]. Additionally, the hematoxylin-eosin
staining results in this study showed that the mucous
epithelium of esophagus, stomach, duodenum, and
rectum did not appear obvious damage [181].
What’s more, the incidence of liver injury in
COVID-19 patients is 39.6% to 43.4%, which is mainly
manifested by increased levels of alanine amino-
transferase (ALT) and aspartate aminotransferase
(AST), as well as hypoalbuminemia [183, 184]. In
contrast, some other studies did not find significant
liver injury in COVID-19 patients [180, 185]. However,
the study also reported that patients with digestive
symptoms had higher mean liver enzyme levels and
longer prothrombin time than those without digestive
symptoms, which could reflect the potential risk of
liver injury [185]. It is hard to explain the variations in
liver test abnormalities among those studies, thus the
effect of SARS-CoV-2 on liver needs further research.
More importantly, another study showed that
patients with gastrointestinal symptoms were
reported more likely to suffer liver injury due to their
elevated ALT and AST compared with patients
without gastrointestinal symptoms [180].
2.8.2. Why gastrointestinal tract occurs?
There are many proposed reasons for the
occurrence of viral gastrointestinal infection. First,
gastrointestinal epithelial cells has been shown to
express ACE2, the receptor of SARS-CoV-2 [181].
Moreover, the positive staining of ACE2 and
SARS-CoV-2 was also observed in gastrointestinal
epithelium from other patients who tested positive for
SARS-CoV-2 RNA in feces [181]. Second, SARS-CoV-2
damages the digestive system through an
inflammatory response. The intestine is the largest
immune organ in the human body, and COVID-19
patients present high inflammatory level [186]. The
chain reaction of inflammatory factors and viremia
may injure the digestive system. Finally, the intestinal
flora plays an important role in body. The number of
intestinal flora in the human intestine is astonishing
and diverse, which is very important for the normal
functioning of digestive system. The virus in the
intestine may cause disorders of intestinal flora,
which result in digestive symptoms [187]. Several
studies have also reported dysbiosis of intestinal flora
in COVID-19 patients [188, 189]. In a study, the author
observed that gut virome and bacteriome in the
COVID-19 patients are notably different from those of
the healthy control, and this difference also exists
between patients of different severity [188]. The study
also confirmed the virome differences and bacteriome
dysbiosis in mouse COVID-19 model. More
importantly, the study reported the differential
expression of immune/infection-related genes in
mouse intestinal epithelial tissues during infection,
such as the polymeric immunoglobulin receptor
(PIGR), interleukin-15 (IL-15), and tribbles
pseudokinase 1 (TRIB1) [188]. Therefore, SARS-CoV-2
may cause the changes in the expression of certain
genes in gastrointestinal tissues, which may be related
to the occurrence of gastrointestinal symptoms. In
conclusion, the correlation between gastrointestinal
symptoms and patients symptoms, diagnosis,
treatment, and outcomes have not been fully
elucidated. It is important and worthy for us to keep
exploring.
3. Treatment of COVID-19
The outbreak of the COVID-19 pandemic has
plunged the world into an unprecedented crisis. The
virus quickly swept across the globe, causing
enormous loss of life, destroying the livelihoods of
billions of people and endangering the global
economy [190]. The cumulative number of confirmed
cases of COVID-19 worldwide has not yet peaked and
the situation remains serious, so the United Nations is
leading and coordinating global efforts to support
countries in their efforts to combat the pandemic.
However, up to now, there is still no good effective
treatment for COVID-19 [191]. Therefore, it has forced
many countries and regions around the world to
quickly carry out the research of the novel
coronavirus.
Particularly, the development of preventive
vaccine against SARS-CoV-2 is the key to control and
prevent the outbreak of a pandemic [192]. According
to the research progress of the COVID-19 vaccine
updated by the World Health Organization, so far
there are 81 new coronavirus vaccines in the clinical
development stage, and more than 180 vaccines are in
the preclinical development stage [193]. Simultaneous
research and development of multiple vaccines will
ensure the quality of the vaccine. The acceleration of
scientific research and clinical trials and the granting
of emergency use authorization by the relevant
government will enable the vaccine to be put on the
market as soon as possible to deal with the sudden
spread of the COVID-19. In addition to vaccines, an
important strategy for the control and treatment of
COVID-19 is to modulate the immune system using
other methods, including the plasma therapy,
suppressing inflammatory cytokines, kinases
inhibitors, cell-based therapies, complement therapy,
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
401
monoclonal antibody therapy and immune
potentiator (Table 1), which are the key
immunotherapeutic approaches to deal with
COVID-19.
3.1. Plasma therapy
It has been demonstrated that convalescent
plasma from COVID-19 patients that have recovered
from the SARS-CoV-2 infection can be utilized as
therapy for patients with COVID-19, without severe
adverse events [194]. Plasma therapy works by
passive transfer antibodies to neutralize the virus.
Clinical data suggested that patients treated with
convalescent plasma had lower mortality than those
who were not [195]. However, this method requires
the collection of plasma from a sufficient number of
convalescent patients. Because of this limitation,
plasma therapy is considered as an option for the
treatment of patients with severe COVID-19.
Importantly, during the progression of the COVID-19
disease, the quality of neutralizing antibodies in
convalescent plasma samples has changed, and
different plasma samples exhibited different antiviral
potentials. Therefore, it is necessary to estimate the
function and titer of neutralizing antibodies from the
donors before treatment [196].
3.2. Cytokine inhibition
The patients with severe COVID-19 were more
likely to generate the cytokine storm, leading to tissue
damage and multi-organ failure. A huge release of
IL-1β, IL-2, IL-6, IL-10, GM-CSF, TNFα, and MCP-1
resulted in immune dysregulation. Therefore, the use
of immunomodulators is beneficial to regulate the
imbalanced immune responses. Current studies
revealed the role of IL-6 in the pathogenesis of
COVID-19, and the development of drugs targeting
the IL-6 pathway is promising for relieving
inflammation in COVID-19 patients [40]. Tocilizumab,
a monoclonal antibody targeting the IL-6 pathway,
has been approved to treat COVID-19. After treatment
with Tocilizumab, the clinical manifestations of
patients have been improved, including rapid control
of fever and improved respiratory function [197].
Similarly, monoclonal antibodies available for
COVID-19 therapy are Anakinra, Etanercept,
Mavrilimumab, and Bevacizumab, which target the
IL-1R, TNFα, GM-CSF, VEGF, respectively, and have
been summarized in Table 1.
3.3. Kinase inhibitors
Janus associated protein kinase (JAK) is a
potential therapeutic target for controlling
SARS-CoV-2 infection. Baricitinib, upadacitinib,
fedratinib, and ruxolitinib, as JAK inhibitors, have
been approved for treating rheumatoid arthritis and
multiple inflammatory diseases. Currently, a
randomized, double-blind, placebo-controlled,
parallel-group Phase III clinical trial of baricitinib in
COVID-19 patients is ongoing, the aim of which is to
investigate whether the baricitinib is effective in
COVID-19 hospitalized patients (NCT04421027).
Bruton tyrosine kinase (BTK) inhibitors are another
group of tyrosine kinase inhibitors. AstraZeneca
initiated a randomized, global clinical trial to evaluate
the efficacy of the acalabrutinib, one of BTK inhibitors,
in the treatment of patients with COVID-19
accompanied by cytokine storm (NCT04497948).
Other kinase inhibitors available for COVID-19
treatment, including sunitinib, a receptor tyrosine
kinase (RTK) inhibitor, and erlotinib, an epidermal
growth factor receptor (EGFR) tyrosine kinase
inhibitor, were shown to block SARS-CoV-2 entry
[198].
3.4. Cell-based therapies
Accumulating evidence has revealed that the
number of NK cells in peripheral blood was
significantly decreased and most of them displayed a
functional exhaustion phenotype in COVID-19
patients. With this in mind, the lack and exhaustion of
NK cells may be one of the reasons for the
unrestricted progression of COVID-19. CYNK-001 is
the allogenic, human placental hematopoietic stem
cell-derived NK cells that can recognize and kill the
virus-infected host cells. A Phase I/II clinical trial is
evaluating its safety, tolerability, and efficacy in
COVID-19 patients (NCT04365101). Another Phase
I/II clinical trial in patients with COVID-19 is
ongoing, which aims to investigate the efficacy of the
constructed NKG2D-ACE2 CAR-NK cells derived
from cord blood in treating severe and critical
COVID-19 (NCT04324996).
Mesenchymal stem cells (MSCs) are a population
of multipotent stem cells with high potential ability of
self-renewal, proliferation, multi-directional differen-
tiation and immunomodulatory [199]. MSCs have an
immunosuppressive effect on different immune cells,
such as T cells, B cells, NK cells and DCs, through
producing a large number of immunosuppressive
agents, including indoleamine-pyrrole 2,
3-dioxygenase (IDO), prostaglandin E2 (PGE2) and
IL-10. Several studies have reported that MSCs served
as a treatment against COVID-19-related cytokine
storm and lung injury. The clinical trials of seven
patients with severe COVID-19 has confirmed the
efficacy and safety of intravenous administration of
MSCs resulting from increasing lymphocyte and
anti-inflammatory cytokines (IL-10) and decreasing
pro-inflammatory cytokines, such as C-reactive
protein (CRP) and TNF [200].
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
402
Table 1. Therapeutic strategies for COVID-19.
Treatment strategy
Agents
Therapeutic target
Plasma Therapy
Convalescent plasma from
COVID-19 patients
Viral proteins
Cytokine therapy
Tocilizumab, Sarilumab, Siltuximab,
Sirukumab, Clazakizumab
IL-6, soluble and membrane
bound IL-6R
Anakinra
IL-1R
Etanercept
TNFα
Mavrilimumab, TJ003234,
Gimsilumab, Lenzilumab
GM-CSFR, GM-CSF
Bevacizumab
VEGF
IFNs prescription
IFN-β-1b, IFN-λ
Kinase inhibitor
Fedratinib, Ruxolitinib, Baricitinib
JAK
Ibrutinib, Acalabrutinib,
Zanubrutinib
BTK
Sunitinib
RTK
Erlotinib
EGFR
Cell-based therapy
NK cells transplantation
NK cells
MSC transplantation
MSC
Tregs adoption
Treg
Monoclonal antibody
therapy
Bamlanivimab
SARS-CoV-2 Spike protein
proliferate
Etesevimab
REGEN-COV
Sotrovimab
4A8
Complement
inhibition
Eculizumab
C5
lung function and lymphocyte recovery
AMY-101
C3
Blood purification
Cytosorb
Cytokines, DAMPs, PAMPs
The balance between effective T cells (Teffs) and
Tregs in the adaptive immune response is most likely
a major factor influencing the outcome of COVID-19
[51, 201]. A clinical study aimed to analyze the global
T cell receptor (TCR) repertoire of peripheral blood
derived Tregs and Teffs, to better understand the
nature of Tregs and Teffs against COVID-19, and to
reveal biomarkers associated with disease severity
(NCT04379466). This is great meaningful for
understanding the pathophysiology of the disease
and designing therapeutics and vaccines.
3.5. Monoclonal antibody therapy
Monoclonal antibodies, which accurately
identify and destroy antigens, play an important role
in disease diagnosis, anti-infection, and anti-tumor.
For viral infections, a neutralizing monoclonal
antibody can specifically neutralize the virus and
prevent the virus from entering the cell to proliferate.
Thus, neutralizing monoclonal antibodies are
regarded as one of the most promising options for the
prevention and treatment of COVID-19. At present,
the SARS-CoV-2 S protein-targeting monoclonal
antibodies are mainly the receptor binding domain
(RBD)-targeting antibodies [202-205]. Bamlanivimab
(LY-CoV555) is a neutralizing monoclonal antibody
that binds to the RBD of SARS-CoV-2 S protein, and
has been shown in Phase II trial to significantly reduce
SARS-CoV-2 levels in patients [202]. According to the
U.S. Food and Drug Administration (FDA)
announcement, Bamlanivimab was the first
monoclonal antibody which receive the emergency
use authorization (EUA) on November 9, 2020.
Etesevimab (LY-CoV016) is another neutralizing
monoclonal antibody that also binds to the RBD of
SARS-CoV-2 S protein [203].The combination therapy
of Bamlanivimab and Etesevimab accelerated the
decline of SARS-CoV-2 viral load and reduced the
mortality rate of COVID-19 patients [203, 206], and
obtained the EUA granted by FDA on February 9,
2021. Unfortunately, whether it was the use of
Bamlanivimab alone or the combined treatment of
Bamlanivimab and Etesevimab, these have been
shown to be unable to resist the mutant virus.
REGEN-COV, combination monoclonal antibodies of
Casirivimab (REGN10933) and Imdevimab
(REGN10987) which bind to the RBD of SARS-CoV-2
S protein [204, 205], has also obtained the EUA from
FDA on November 21, 2020, and more importantly,
REGEN-COV still retains its effectiveness against a
variety of mutant viruses [205]. Sotrovimab
(VIR-7831), a monoclonal antibody which also binds
to the RBD of SARS-CoV-2 S protein has shown that
the risk of hospitalization or death in the Sotrovimab
group was 85% lower than that of the control group
[207], and Sotrovimab obtained the EUA from FDA on
May 26, 2021. Additionally, 4A8, as an N-terminal
domain (NTD)-targeting antibody, has a strong virus
neutralization ability [208]. Combination of the
NTD-targeting antibody with RBD-targeting antibody
may avoid the escaping mutations of the virus and
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
403
serve as promising “cocktail” therapeutics. Of course,
there are many other monoclonal antibodies for the
treatment of COVID-19 in the research and
development stage or clinical trial stage, and it is of
great significance and contribution to the fight against
the global epidemic.
3.6. Other therapies
An adjunctive therapy available for COVID-19 is
cytosorb, which absorbs a broad spectrum of
cytokines, damage-associated molecular patterns
(DAMPs), and pathogen-associated molecular
patterns (PAMPs) in the blood circulation to reduce
inflammation and improve immunopathology of the
disease [209]. Complement inhibitors have emerged
as the drug candidates against SARS-CoV-2 infection.
In a cohort study, COVID-19 patients were treated
with the eulizumab and the cyclic peptide AMY-101
to block complement C5 and C3, respectively, and it
was shown that complement inhibition alleviated
hyper-inflammation characterized by a significant
decrease in serum IL-6 and CRP, reduced neutrophil
counts, and markedly improved lung function and
lymphocyte recovery [210]. Because the surface
ligands of SARS-CoV-2 are constantly altered to
escape neutralizing antibodies, one strategy is to
apply drugs that block receptors for such ligands on
host cells, such as ACE2. Immune potentiator
treatment strategies aim to stimulate innate and
adaptive immunity through multiple mechanisms to
eliminate viral infections. These potentiators include
antimicrobial peptides, immune checkpoint
inhibitors, pattern recognition receptor (PRR) ligands,
and signaling compartments [211]. Neutralizing
antibodies against PD-1, alone or in combination with
thymosin, are under investigation for their efficacy in
COVID-19 cases (NCT04268537). Additionally, the
use of corticosteroid to treat patients with COVID-19
remains controversial currently, and relevant clinical
trials are ongoing (NCT04244591).
4. Conclusions and Perspectives
Similar to SARS, COVID-19 manifests mainly as
the symptoms of respiratory system, but emerging
evidences as mentioned above suggest that
SARS-CoV-2 affects various other systems in humans
as well. Clinical manifestations of multisystem
infection are more unpredictable to thus make the
treatment of COVID-19 more difficult. Therefore, it is
necessary to perform a comprehensive physical
assessment and provide a systematic therapeutic
schedule for each inpatient, maybe with different
symptoms. The review provides a novel perspective
on COVID-19 from the infection with multisystem
involvement to help the health and medical
community to acquire available information.
Additionally, scientific researches still need to be
funded and executed to reveal more details about the
molecular mechanism of SARS-CoV-2 to solve the
following three solemn problems: (1) How to
diagnose accurately as early as possible? (2) How to
effectively control the spread of the virus? (3) How to
cure the COVID-19 efficiently?
Acknowledgments
Funding
The work was supported by the National
Natural Science Foundation of China (62005085), and
the China Postdoctoral Science Foundation
(2021M691093).
Author Contributions
QS, JL, HCC, ZZ, SG and QHW conceived and
drafted the manuscript. HCC, QS, ZZ and SG drew
the figures. HCC discussed the concepts of the
manuscript. XRA checked the manuscript.
Competing Interests
The authors have declared that no competing
interest exists.
References
1. Khailany RA, Safdar M, Ozaslan M. Genomic characterization of a novel
SARS-CoV-2. Gene Rep. 2020; 19: 100682.
2. Finkel Y, Mizrahi O, Nachshon A, Weingarten-Gabbay S, Morgenstern
D, Yahalom-Ronen Y, et al. The coding capacity of SARS-CoV-2. Nature.
2021; 589: 125-130.
3. Lei XB, Dong XJ, Ma RY, Wang WJ, Xiao X, Tian ZQ, et al. Activation and
evasion of type I interferon responses by SARS-CoV-2. Nat Commun.
2020; 11: 3810.
4. Xia HJ, Cao ZG, Xie XP, Zhang XW, Chen JYC, Wang HL, et al. Evasion
of type I Interferon by SARS-CoV-2. Cell Rep. 2020; 33: 108234.
5. Zhang Y, Xiao M, Zhang S, Xia P, Cao W, Jiang W, et al. Coagulopathy
and antiphospholipid antibodies in patients with Covid-19. N Engl J
Med. 2020; 382: e38.
6. Xu H, Zhong L, Deng J, Peng J, Dan H, Zeng X, et al. High expression of
ACE2 receptor of 2019-nCoV on the epithelial cells of oral mucosa. Int J
Oral Sci. 2020; 12: 8.
7. Ziegler CGK, Allon SJ, Nyquist SK, Mbano IM, Miao VN, Tzouanas CN,
et al. SARS-CoV-2 receptor ACE2 is an interferon-stimulated gene in
human airway epithelial cells and is detected in specific cell subsets
across tissues. Cell. 2020; 181: 1016-35 e19.
8. Adil M, Verma A, Rudraraju M, Narayanan S, Somanath P.
Akt-independent effects of triciribine on ACE2 expression in human
lung epithelial cells: Potential benefits in restricting SARS-CoV2
infection. J Cell Physiol. 2021; 10: 1002.
9. Wang J, Zhao S, Liu M, Zhao Z, Zhang Y. ACE2 expression by colonic
epithelial cells is associated with viral infection, immunity and energy
metabolism. medRxiv. 2020.
10. José S, Jan W, Daniel B. ACE2 alterations in kidney disease. Nephrol Dial
Transplant 2013; 46: 1339-48.
11. Song H, Seddighzadeh B, Cooperberg MR, Huang FW. Expression of
ACE2, the SARS-CoV-2 receptor, and TMPRSS2 in prostate epithelial
cells. Eur Urol. 2020; 78: 296-8.
12. Ghavami S, Yeganeh B, Zeki AA, Shojaei S, Kenyon NJ, Ott S, et al.
Autophagy and the unfolded protein response promote pro-fibrotic
effects of TGFβ1 in human lung fibroblasts. Am J Physiol Lung Cell Mol
Physiol. 2017; 314: L493-L504.
13. Schmidinger G, Hanselmayer G, Pieh S, Lackner B, Skorpik C. Effect of
tenascin and fibronectin on the migration of human corneal fibroblasts. J
Cataract Refract Surg. 2003; 29: 354-60.
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
404
14. Har A, Snbbc D. The epidemiology and pathogenesis of coronavirus
disease (COVID-19) outbreak. J Autoimmun. 2020; 109: 102433.
15. Xiong Y, Liu Y, Cao L, Wang D, Guo M, Jiang A, et al. Transcriptomic
characteristics of bronchoalveolar lavage fluid and peripheral blood
mononuclear cells in COVID-19 patients. Emerg Microbes Infect. 2020; 9:
761-70.
16. Lechowicz K, Drodal S, Machaj F, Rosik J, Kotfis K. COVID-19: The
potential treatment of pulmonary fibrosis associated with SARS-CoV-2
infection. J Clin Med. 2020; 9: 1917.
17. Selman M, King T, Parodo A. Idiopathic pulmonary fibrosis: prevailing
and evolving hypotheses about its pathogenesis and implications for
therapy. Ann Intern Med. 2001; 134: 136-51.
18. Zhou Z, Ren L, Zhang L, Zhong J, Xiao Y, Jia Z, et al. Heightened innate
immune responses in the respiratory tract of COVID-19 patients. Cell
Host Microbe. 2020; 27: 883-90 e2.
19. Blanco-Melo D, Nilsson-Payant BE, Liu WC, Uhl S, Hoagland D, Moller
R, et al. Imbalanced host response to SARS-CoV-2 drives development of
COVID-19. Cell. 2020; 181: 1036-45 e9.
20. Sette A, Crotty S. Adaptive immunity to SARS-CoV-2 and COVID-19.
Cell. 2021; 184: 861-80.
21. Huang CL, Wang YM, Li XW, Ren LL, Zhao JP, Hu Y, et al. Clinical
features of patients infected with 2019 novel coronavirus in Wuhan,
China. Lancet. 2020; 395: 497-506.
22. Shi SB, Qin M, Shen B, Cai YL, Liu T, Yang F, et al. Association of cardiac
injury with mortality in hospitalized patients with COVID-19 in Wuhan,
China. Jama Cardiol. 2020; 5: 802-10.
23. Lucas C, Wong P, Klein J, Castro TBR, Silva J, Sundaram M, et al.
Longitudinal analyses reveal immunological misfiring in severe
COVID-19. Nature. 2020; 584: 463-9.
24. Wilk AJ, Rustagi A, Zhao NQ, Roque J, Martinez-Colon GJ, McKechnie
JL, et al. A single-cell atlas of the peripheral immune response in patients
with severe COVID-19. Nat Med. 2020; 26: 1070-6.
25. Schulte-Schrepping J, Reusch N, Paclik D, Bassler K, Schlickeiser S,
Zhang B, et al. Severe COVID-19 is marked by a dysregulated myeloid
cell compartment. Cell. 2020; 182: 1419-40 e23.
26. Guo C, Li B, Ma H, Wang X, Cai P, Yu Q, et al. Single-cell analysis of two
severe COVID-19 patients reveals a monocyte-associated and
tocilizumab-responding cytokine storm. Nat Commun. 2020; 11: 3924.
27. Wen W, Su W, Tang H, Le W, Zhang X, Zheng Y, et al. Immune cell
profiling of COVID-19 patients in the recovery stage by single-cell
sequencing. Cell Discov. 2020; 6: 31.
28. Liao M, Liu Y, Yuan J, Wen Y, Xu G, Zhao J, et al. Single-cell landscape of
bronchoalveolar immune cells in patients with COVID-19. Nat Med.
2020; 26: 842-4.
29. Chua RL, Lukassen S, Trump S, Hennig BP, Wendisch D, Pott F, et al.
COVID-19 severity correlates with airway epithelium-immune cell
interactions identified by single-cell analysis. Nat Biotechnol. 2020; 38:
970-9.
30. Sanchez-Cerrillo I, Landete P, Aldave B, Sanchez-Alonso S,
Sanchez-Azofra A, Marcos-Jimenez A, et al. Differential redistribution of
activated monocyte and dendritic cell subsets to the lung associates with
severity of COVID-19. medRxiv. 2020.
31. Zheng M, Gao Y, Wang G, Song G, Liu S, Sun D, et al. Functional
exhaustion of antiviral lymphocytes in COVID-19 patients. Cell Mol
Immunol. 2020; 17: 533-5.
32. Robbiani DF, Gaebler C, Muecksch F, Lorenzi JCC, Wang Z, Cho A, et al.
Convergent antibody responses to SARS-CoV-2 in convalescent
individuals. Nature. 2020; 584: 437-42.
33. Maucourant C, Filipovic I, Ponzetta A, Aleman S, Cornillet M, Hertwig
L, et al. Natural killer cell immunotypes related to COVID-19 disease
severity. Sci Immunol. 2020; 5: eabd6832.
34. Su Y, Chen D, Yuan D, Lausted C, Choi J, Dai CL, et al. Multi-omics
resolves a sharp disease-state shift between mild and moderate
COVID-19. Cell. 2020; 183: 1479-95 e20.
35. Rodriguez L, Pekkarinen PT, Lakshmikanth T, Tan Z, Consiglio CR, Pou
C, et al. Systems-level immunomonitoring from acute to recovery phase
of severe COVID-19. Cell Rep Med. 2020; 1: 100078.
36. de Nooijer AH, Grondman I, Janssen NAF, Netea MG, Willems L, van de
Veerdonk FL, et al. Complement activation in the disease course of
coronavirus disease 2019 and its effects on clinical outcomes. J Infect Dis.
2021; 223: 214-24.
37. Holter JC, Pischke SE, de Boer E, Lind A, Jenum S, Holten AR, et al.
Systemic complement activation is associated with respiratory failure in
COVID-19 hospitalized patients. Proc Natl Acad Sci U S A. 2020; 117:
25018-25.
38. He CT, Qin M, Sun X. Highly pathogenic coronaviruses: thrusting
vaccine development in the spotlight. Acta Pharmacol Sin B. 2020; 10:
1175-91.
39. Shen B, Yi X, Sun Y, Bi X, Du J, Zhang C, et al. Proteomic and
metabolomic characterization of COVID-19 patient sera. Cell. 2020; 182:
59-72 e15.
40. Giamarellos-Bourboulis EJ, Netea MG, Rovina N, Akinosoglou K,
Antoniadou A, Antonakos N, et al. Complex immune dysregulation in
COVID-19 patients with severe respiratory failure. Cell Host Microbe.
2020; 27: 992-1000 e3.
41. Hadjadj J, Yatim N, Barnabei L, Corneau A, Boussier J, Smith N, et al.
Impaired type I interferon activity and inflammatory responses in severe
COVID-19 patients. Science. 2020; 369: 718.
42. Lee JS, Park S, Jeong HW, Ahn JY, Choi SJ, Lee H, et al.
Immunophenotyping of COVID-19 and influenza highlights the role of
type I interferons in development of severe COVID-19. Sci Immunol.
2020; 5: eabd1554.
43. Arunachalam PS, Wimmers F, Mok CKP, Perera R, Scott M, Hagan T, et
al. Systems biological assessment of immunity to mild versus severe
COVID-19 infection in humans. Science. 2020; 369: 1210-20.
44. Tan L, Wang Q, Zhang D, Ding J, Huang Q, Tang YQ, et al.
Lymphopenia predicts disease severity of COVID-19: a descriptive and
predictive study. Signal Transduct Target Ther. 2020; 5: 33.
45. Grifoni A, Weiskopf D, Ramirez SI, Mateus J, Dan JM, Moderbacher CR,
et al. Targets of T cell responses to SARS-CoV-2 coronavirus in humans
with COVID-19 disease and unexposed individuals. Cell. 2020; 181:
1489-501 e15.
46. Rydyznski Moderbacher C, Ramirez SI, Dan JM, Grifoni A, Hastie KM,
Weiskopf D, et al. Antigen-specific adaptive immunity to SARS-CoV-2 in
acute COVID-19 and associations with age and disease severity. Cell.
2020; 183: 996-1012 e19.
47. Tan AT, Linster M, Tan CW, Le Bert N, Chia WN, Kunasegaran K, et al.
Early induction of functional SARS-CoV-2-specific T cells associates with
rapid viral clearance and mild disease in COVID-19 patients. Cell Rep.
2021; 34: 108728.
48. Bacher P, Rosati E, Esser D, Martini GR, Saggau C, Schiminsky E, et al.
Low-avidity CD4(+) T cell responses to SARS-CoV-2 in unexposed
individuals and humans with severe COVID-19. Immunity. 2020; 53:
1258-71 e5.
49. Braun J, Loyal L, Frentsch M, Wendisch D, Georg P, Kurth F, et al.
SARS-CoV-2-reactive T cells in healthy donors and patients with
COVID-19. Nature. 2020; 587: 270-4.
50. Meckiff BJ, Ramirez-Suastegui C, Fajardo V, Chee SJ, Kusnadi A, Simon
H, et al. Single-cell transcriptomic analysis of SARS-CoV-2 reactive CD4
(+) T cells. bioRxiv. 2020.
51. Meckiff BJ, Ramirez-Suastegui C, Fajardo V, Chee SJ, Kusnadi A, Simon
H, et al. Imbalance of regulatory and cytotoxic SARS-CoV-2-reactive
CD4(+) T cells in COVID-19. Cell. 2020; 183: 1340-53 e16.
52. Schulien I, Kemming J, Oberhardt V, Wild K, Seidel LM, Killmer S, et al.
Characterization of pre-existing and induced SARS-CoV-2-specific
CD8(+) T cells. Nat Med. 2021; 27: 78-85.
53. Weiskopf D, Schmitz KS, Raadsen MP, Grifoni A, Okba NMA, Endeman
H, et al. Phenotype and kinetics of SARS-CoV-2-specific T cells in
COVID-19 patients with acute respiratory distress syndrome. Sci
Immunol. 2020; 5: eabd2071.
54. Sekine T, Perez-Potti A, Rivera-Ballesteros O, Stralin K, Gorin JB, Olsson
A, et al. Robust T cell immunity in convalescent individuals with
asymptomatic or mild COVID-19. Cell. 2020; 183: 158-68 e14.
55. Riddell SR, Greenberg PD. The use of anti-CD3 and anti-CD28
monoclonal antibodies to clone and expand human antigen-specific T
cells. J Immunol Methods. 1990; 128: 189-201.
56. Chang HC, Zou ZZ, Li J, Shen Q, Liu L, An XR, et al. Photoactivation of
mitochondrial reactive oxygen species-mediated Src and protein kinase
C pathway enhances MHC class II-restricted T cell immunity to tumours.
Cancer Lett. 2021; 523: 57-71.
57. Long QX, Liu BZ, Deng HJ, Wu GC, Deng K, Chen YK, et al. Antibody
responses to SARS-CoV-2 in patients with COVID-19. Nat Med. 2020; 26:
845-8.
58. Ripperger TJ, Uhrlaub JL, Watanabe M, Wong R, Castaneda Y, Pizzato
HA, et al. Orthogonal SARS-CoV-2 serological assays enable surveillance
of low-prevalence communities and reveal durable humoral immunity.
Immunity. 2020; 53: 925-33 e4.
59. Zhao JJ, Yuan Q, Wang HY, Liu W, Liao XJ, Su YY, et al. Antibody
responses to SARS-CoV-2 in patients with novel coronavirus disease
2019. Clin Infect Dis. 2020; 71: 2027-34.
60. Piccoli L, Park YJ, Tortorici MA, Czudnochowski N, Walls AC,
Beltramello M, et al. Mapping neutralizing and immunodominant sites
on the SARS-CoV-2 spike receptor-binding domain by structure-guided
high-resolution serology. Cell. 2020; 183: 1024-42 e21.
61. Liu L, Wang P, Nair MS, Yu J, Rapp M, Wang Q, et al. Potent
neutralizing monoclonal antibodies directed to multiple epitopes on the
SARS-CoV-2 spike. bioRxiv. 2020.
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
405
62. Wajnberg A, Amanat F, Firpo A, Altman DR, Bailey MJ, Mansour M, et
al. Robust neutralizing antibodies to SARS-CoV-2 infection persist for
months. Science. 2020; 370: 1227-30.
63. Wang X, Guo X, Xin Q, Pan Y, Hu Y, Li J, et al. Neutralizing antibody
responses to severe acute respiratory syndrome coronavirus 2 in
coronavirus disease 2019 inpatients and convalescent patients. Clin
Infect Dis. 2020; 71: 2688-94.
64. Dan JM, Mateus J, Kato Y, Hastie KM, Yu ED, Faliti CE, et al.
Immunological memory to SARS-CoV-2 assessed for up to 8 months
after infection. Science. 2021; 371: 587.
65. Gaebler C, Wang Z, Lorenzi JCC, Muecksch F, Finkin S, Tokuyama M, et
al. Evolution of antibody immunity to SARS-CoV-2. bioRxiv. 2020.
66. Zuo JM, Dowell AC, Pearce H, Verma K, Long HM, Begum J, et al.
Robust SARS-CoV-2-specific T cell immunity is maintained at 6 months
following primary infection. Nat Immunol. 2021; 22: 620-6.
67. Eastin C, Eastin T. Clinical characteristics of coronavirus disease 2019 in
China. J Emerg Med. 2020; 58: 711-2.
68. Liao YC, Liang WG, Chen FW, Hsu JH, Yang JJ, Chang MS. IL-19 induces
production of IL-6 and TNF-alpha and results in cell apoptosis through
TNF-alpha. J Immunol. 2002; 169: 4288-97.
69. Aggarwal S, Gollapudi S, Gupta S. Increased TNF-alpha-induced
apoptosis in lymphocytes from aged humans: changes in TNF-alpha
receptor expression and activation of caspases. J Immunol. 1999; 162:
2154-61.
70. Singh S, Sharma A, Arora SK. High producer haplotype (CAG) of
-863C/A, -308G/A and -238G/A polymorphisms in the promoter region
of TNF-α gene associate with enhanced apoptosis of lymphocytes in
HIV-1 subtype C infected individuals from North India. PloS one. 2014;
9: e98020.
71. Chan FW, Zhang AJ, Yuan S, Kwok-Man PV, Chan CS, Chak-Yiu LA, et
al. Simulation of the clinical and pathological manifestations of
Coronavirus Disease 2019 (COVID-19) in golden Syrian hamster model:
implications for disease pathogenesis and transmissibility. Clin Infect
Dis. 2020; 71: 2428-46.
72. Driggin E, Madhavan MV, Bikdeli B, Chuich T, Laracy J, Biondi-Zoccai
G, et al. Cardiovascular considerations for patients, health care workers,
and health systems during the COVID-19 pandemic. J Am Coll Cardiol.
2020; 75: 2352-71.
73. Wang D, Hu B, Hu C, Zhu F, Liu X, Zhang J, et al. Clinical characteristics
of 138 hospitalized patients with 2019 novel coronavirus-infected
pneumonia in Wuhan, China. Jama. 2020; 323: 1061-9.
74. Tadic M, Saeed S, Grassi G, Taddei S, Mancia G, Cuspidi C.
Hypertension and COVID-19: ongoing controversies. Front Cardiovasc
Med. 2021; 8: 639222.
75. Madjid M, Safavi-Naeini P, Solomon SD, Vardeny O. Potential effects of
coronaviruses on the cardiovascular system a review. Jama Cardiol.
2020; 5: 831-40.
76. Song J, Deng Y-K, Wang H, Wang Z-C, Liao B, Ma J, et al. Self-reported
taste and smell disorders in patients with COVID-19: distinct features in
China. Curr Med Sci. 2020; 41: 14-23.
77. Zheng YY, Ma YT, Zhang JY, Xie X. COVID-19 and the cardiovascular
system. Nat Rev Cardiol. 2020; 17: 259-60.
78. Arachchillage DR, Laffan M. Abnormal coagulation parameters are
associated with poor prognosis in patients with novel coronavirus
pneumonia. J Thromb Haemost. 2020; 18: 844-7.
79. Lew, Thomas WK. Acute respiratory distress syndrome in critically ill
patients with severe acute respiratory syndrome. Jama. 2003; 290: 374-80.
80. Madjid M, Aboshady I, Awan I, Litovsky S, Casscells SW. Influenza and
cardiovascular disease: is there a causal relationship? Tex Heart Inst J.
2004; 31: 4-13.
81. Bergmann CC, Lane Te Fau - Stohlman SA, Stohlman SA. Coronavirus
infection of the central nervous system: host-virus stand-off. Nat Rev
Microbiol. 2006; 4: 121-32.
82. Song E, Zhang C, Israelow B, Lu P, Weizman O-E, Liu F, et al.
Neuroinvasive potential of SARS-CoV-2 revealed in a human brain
organoid model. bioRxiv 2020.
83. Yang L, Han Y, Nilsson-Payant BE, Gupta V, Wang P, Duan X, et al. A
human pluripotent stem cell-based platform to study SARS-CoV-2
tropism and model virus infection in human cells and organoids. Cell
Stem Cell. 2020; 27: 125-36 e7.
84. Solomon IH, Normandin E, Bhattacharyya S, Mukerji SS, Keller K, Ali
AS, et al. Neuropathological features of Covid-19. N Engl J Med. 2020;
383: 989-92.
85. Ye M, Ren Y, Lv T. Encephalitis as a clinical manifestation of COVID-19.
Brain Behav Immun. 2020; 88: 945-6.
86. Bernard-Valnet R, Pizzarotti B, Anichini A, Demars Y, Russo E,
Schmidhauser M, et al. Two patients with acute meningoencephalitis
concomitant with SARS-CoV-2 infection. Eur J Neurol. 2020; 27: e43-e4.
87. Finsterer J, Stollberger C. Update on the neurology of COVID-19. J Med
Virol. 2020; 92: 2316-8.
88. Baig AM, Khaleeq A, Ali U, Syeda H. Evidence of the COVID-19 virus
targeting the CNS: tissue distribution, host-virus interaction, and
proposed neurotropic mechanisms. ACS Chem Neurosci. 2020; 11: 995-8.
89. Kempuraj D, Selvakumar GP, Ahmed ME, Raikwar SP, Thangavel R,
Khan A, et al. COVID-19, mast cells, cytokine storm, psychological
stress, and neuroinflammation. Neuroscientist. 2020; 26: 402-14.
90. Brann DH, Tsukahara T, Weinreb C, Lipovsek M, Van den Berge K,
Gong B, et al. Non-neuronal expression of SARS-CoV-2 entry genes in
the olfactory system suggests mechanisms underlying
COVID-19-associated anosmia. Sci Adv. 2020; 6: eabc5801.
91. Gleeson J, Wang L, Sievert D, Clark A, Federman H, Gastfriend B, et al. A
human 3D neural assembloid model for SARS-CoV-2 infection. Research
square. 2021: 1-16.
92. Solomon T. Neurological infection with SARS-CoV-2 - the story so far.
Nat Rev Neurol. 2021; 17: 65-6.
93. Andrews MG, Mukhtar T, Eze UC, Simoneau CR, Perez Y, Mostajo-Radji
MA, et al. Tropism of SARS-CoV-2 for developing human cortical
astrocytes. bioRxiv. 2021.
94. Crunfli F, Carregari VC, Veras FP, Vendramini PH, Valença AGF,
Antunes ASLM, et al. SARS-CoV-2 infects brain astrocytes of COVID-19
patients and impairs neuronal viability. medRxiv. 2021.
95. Mao L, Wang M, Shengcai C, He Q, Chang J, Hong C, et al. Neurological
Manifestations of Hospitalized Patients with COVID-19 in Wuhan,
China: A Retrospective Case Series Study. medRxiv. 2020.
96. Oxley TJ, Mocco J, Majidi S, Kellner CP, Shoirah H, Singh IP, et al.
Large-vessel stroke as a presenting feature of Covid-19 in the Young. N
Engl J Med. 2020; 382: e60.
97. Merkler A, Parikh N, Mir S, Gupta A, Kamel H, Lin E, et al. Risk of
ischemic stroke in patients with coronavirus disease 2019 (COVID-19) vs
patients with influenza. JAMA Neurol. 2020; 77: 1-7.
98. Chen T, Wu D, Chen H, Yan W, Yang D, Chen G, et al. Clinical
characteristics of 113 deceased patients with coronavirus disease 2019:
retrospective study. BMJ. 2020; 368: m1091.
99. Mao L, Jin H, Wang M, Hu Y, Chen S, He Q, et al. Neurologic
manifestations of hospitalized patients with coronavirus disease 2019 in
Wuhan, China. JAMA Neurol. 2020; 77: 683-90.
100. Belvis R. Headaches during COVID-19: my clinical case and review of
the literature. Headache. 2020; 60: 1422-6.
101. Jones VG, Mills M, Suarez D, Hogan CA, Yeh D, Segal JB, et al.
COVID-19 and Kawasaki disease: novel virus and novel case. Hosp
Pediatr. 2020; 10: 537-40.
102. Verdoni L, Mazza A, Gervasoni A, Martelli L, Ruggeri M, Ciuffreda M, et
al. An outbreak of severe Kawasaki-like disease at the Italian epicentre of
the SARS-CoV-2 epidemic: an observational cohort study. Lancet. 2020;
395: 1771-8.
103. Carfì A, Bernabei R, Landi F. Persistent symptoms in patients after acute
COVID-19. Jama. 2020; 324: 603-5.
104. Lubbe L, Cozier GE, Oosthuizen D, Acharya KR, Sturrock ED. ACE2 and
ACE: structure-based insights into mechanism, regulation and receptor
recognition by SARS-CoV. Clin Sci (Lond). 2020; 134: 2851-71.
105. Divani AA, Andalib S, Di Napoli M, Lattanzi S, Hussain MS, Biller J, et
al. Coronavirus disease 2019 and stroke: clinical manifestations and
pathophysiological insights. J Stroke Cerebrovasc Dis. 2020; 29: 104941.
106. Inciardi RM, Adamo M, Lupi L, Cani DS, Di Pasquale M, Tomasoni D, et
al. Characteristics and outcomes of patients hospitalized for COVID-19
and cardiac disease in Northern Italy. Eur Heart J. 2020; 41: 1821-9.
107. Goyal P, Choi JJ, Pinheiro LC, Schenck EJ, Chen R, Jabri A, et al. Clinical
characteristics of Covid-19 in New York city. N Engl J Med. 2020; 382:
2372-4.
108. Coolen T, Lolli V, Sadeghi N, Rovai A, Trotta N, Taccone FS, et al. Early
postmortem brain MRI findings in COVID-19 non-survivors. Neurology.
2020; 95: e2016-e27.
109. Elkind MSV, Boehme AK, Smith CJ, Meisel A, Buckwalter MS. Infection
as a stroke risk factor and determinant of outcome after stroke. Stroke.
2020; 51: 3156-68.
110. Pal R, Banerjee M. COVID-19 and the endocrine system: exploring the
unexplored. J Endocrinol Invest. 2020; 43: 1027-31.
111. Marazuela M, Giustina A, Puig-Domingo M. Endocrine and metabolic
aspects of the COVID-19 pandemic. Rev Endocr Metab Disord. 2020, 21:
495-507.
112. Carsana L, Sonzogni A, Nasr A, Rossi RS, Pellegrinelli A, Zerbi P, et al.
Pulmonary post-mortem findings in a series of COVID-19 cases from
northern Italy: a two-centre descriptive study. Lancet Infect Dis. 2020; 20:
1135-40.
113. Bost P, Giladi A, Liu Y, Bendjelal Y, Xu G, David E, et al. Host-viral
infection maps reveal signatures of severe COVID-19 patients. Cell. 2020;
181: 1475-88.e12.
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
406
114. Grasselli G, Zangrillo A, Zanella A, Antonelli M, Cabrini L, Castelli A, et
al. Baseline characteristics and outcomes of 1591 patients infected with
SARS-CoV-2 admitted to ICUs of the lombardy region, Italy. Jama. 2020;
323: 1574-81.
115. Reichard RR, Kashani KB, Boire NA, Constantopoulos E, Guo Y,
Lucchinetti CF. Neuropathology of COVID-19: a spectrum of vascular
and acute disseminated encephalomyelitis (ADEM)-like pathology. Acta
Neuropathol. 2020; 140: 1-6.
116. Kim JE, Heo JH, Kim HO, Song SH, Park SS, Park TH, et al. Neurological
complications during treatment of Middle East Respiratory Syndrome. J
Clin Neurol. 2017; 13: 227-33.
117. Camdessanche JP, Morel J, Pozzetto B, Paul S, Tholance Y,
Botelho-Nevers E. COVID-19 may induce Guillain-Barré syndrome. Rev
Neurol (Paris). 2020; 176: 516-8.
118. Toscano G, Palmerini F, Ravaglia S, Ruiz L, Invernizzi P, Cuzzoni MG, et
al. Guillain-Barré Syndrome associated with SARS-CoV-2. N Engl J Med.
2020; 382: 2574-6.
119. Zhao H, Shen D, Zhou H, Liu J, Chen S. Guillain-Barré syndrome
associated with SARS-CoV-2 infection: causality or coincidence? Lancet
Neurol. 2020; 19: 383-4.
120. Abdelnour L, Eltahir Abdalla M, Babiker S. COVID 19 infection
presenting as motor peripheral neuropathy. J Formos Med Assoc. 2020;
119: 1119-20.
121. Coen M, Jeanson G, Culebras Almeida LA, Hübers A, Stierlin F, Najjar I,
et al. Guillain-Barré syndrome as a complication of SARS-CoV-2
infection. Brain Behav Immun. 2020; 87: 111-2.
122. Katyal N, Narula N, Acharya S, Govindarajan R. Neuromuscular
complications with SARS-COV-2 infection: A review. Front Neurol.
2020; 11: 1052.
123. Ghosh R, Roy D, Sengupta S, Benito-León J. Autonomic dysfunction
heralding acute motor axonal neuropathy in COVID-19. J Neurovirol.
2020; 26: 964-6.
124. Wei H, Yin H, Huang M, Guo Z. The 2019 novel cornoavirus pneumonia
with onset of oculomotor nerve palsy: a case study. J Neurol. 2020; 267:
1550-3.
125. Wan Y, Cao S, Fang Q, Wang M, Huang Y. Coronavirus disease 2019
complicated with Bell’s palsy: a case report. Research Square. 2020.
126. Rajan S, Kaas B, Moukheiber E. Movement disorders emergencies. Semin
Neurol. 2019; 39: 125-36.
127. Valiuddin HM, Kalajdzic A, Rosati J, Boehm K, Hill D. Update on
Neurological manifestations of SARS-CoV-2. West J Emerg Med. 2020;
21: 45-51.
128. Brown EG, Chahine LM, Goldman SM, Korell M, Mann E, Kinel DR, et
al. The effect of the COVID-19 pandemic on people with Parkinson's
disease. J Parkinsons Dis. 2020; 10: 1365-77.
129. Dhar N, Dhar S, Timar R, Lucas S, Lamb LE, Chancellor MB. De Novo
urinary symptoms associated with COVID-19: COVID-19-associated
cystitis. J Clin Med Res. 2020; 12: 681-2.
130. Mumm JN, Osterman A, Ruzicka M, Stihl C, Vilsmaier T, Munker D, et
al. Urinary frequency as a possibly overlooked symptom in COVID-19
patients: does SARS-CoV-2 cause viral cystitis? Eur Urol. 2020; 78: 624-8.
131. La Vignera S, Cannarella R, Condorelli RA, Torre F, Aversa A, Calogero
AE. Sex-specific SARS-CoV-2 mortality: among hormone-modulated
ACE2 expression, risk of venous thromboembolism and
hypovitaminosis D. Int J Mol Sci. 2020; 21: 2948.
132. Karabulut I, Cinislioglu AE, Cinislioglu N, Yilmazel FK, Utlu M, Alay H,
et al. The effect of the presence of lower urinary system symptoms on the
prognosis of COVID-19: preliminary results of a prospective study. Urol
Int. 2020; 104: 853-8.
133. Borges do Nascimento IJ, Cacic N, Abdulazeem HM, von Groote TC,
Jayarajah U, Weerasekara I, et al. Novel coronavirus infection
(COVID-19) in humans: a scoping review and meta-analysis. J Clin Med.
2020; 9: 941.
134. Channappanavar R, Fett C, Mack M, Ten Eyck PP, Meyerholz DK,
Perlman S. Sex-based differences in susceptibility to severe acute
respiratory syndrome coronavirus infection. J Immunol. 2017; 198:
4046-53.
135. Mcnicholas T, Kirby R. Benign prostatic hyperplasia and male lower
urinary tract symptoms. Am Fam Physician. 2012; 86: 359.
136. Thorpe A, Neal D. Benign prostatic hyperplasia. Lancet. 2003; 361:
1359-67.
137. Peerapornratana S, Manrique-Caballero CL, Gómez H, Kellum JA. Acute
kidney injury from sepsis: current concepts, epidemiology,
pathophysiology, prevention and treatment. Kidney Int. 2019; 96:
1083-99.
138. Gabarre P, Dumas G, Dupont T, Darmon M, Azoulay E, Zafrani L. Acute
kidney injury in critically ill patients with COVID-19. Intensive Care
Med. 2020; 46: 1339-48.
139. Yang X, Yu Y, Xu J, Shu H, Xia J, Liu H, et al. Clinical course and
outcomes of critically ill patients with SARS-CoV-2 pneumonia in
Wuhan, China: a single-centered, retrospective, observational study.
Lancet Respir Med. 2020; 8: 475-81.
140. Zhou F, Yu T, Du R, Fan G, Liu Y, Liu Z, et al. Clinical course and risk
factors for mortality of adult inpatients with COVID-19 in Wuhan,
China: a retrospective cohort study. Lancet. 2020; 395: 1054-62.
141. Varga Z, Flammer AJ, Steiger P, Haberecker M, Andermatt R,
Zinkernagel AS, et al. Endothelial cell infection and endotheliitis in
COVID-19. Lancet. 2020; 395: 1417-8.
142. Husain-Syed F, Wilhelm J, Kassoumeh S, Birk HW, Herold S, Vadász I, et
al. Acute kidney injury and urinary biomarkers in hospitalized patients
with coronavirus disease-2019. Nephrol Dial Transplant. 2020; 35:
1271-4.
143. Wang N, Qin L, Ma L, Yan H. Effect of severe acute respiratory
syndrome coronavirus-2 (SARS-CoV-2) on reproductive system. Stem
Cell Res. 2021; 52: 102189.
144. Reis FM, Bouissou DR, Pereira VM, Camargos AF, dos Reis AM, Santos
RA. Angiotensin-(1-7), its receptor Mas, and the angiotensin-converting
enzyme type 2 are expressed in the human ovary. Fertil Steril. 2011; 95:
176-81.
145. Vaz-Silva J, Carneiro MM, Ferreira MC, Pinheiro SV, Silva DA,
Silva-Filho AL, et al. The vasoactive peptide angiotensin-(1-7), its
receptor Mas and the angiotensin-converting enzyme type 2 are
expressed in the human endometrium. Reprod Sci. 2009; 16: 247-56.
146. Douglas GC, O'Bryan MK, Hedger MP, Lee DK, Yarski MA, Smith AI, et
al. The novel angiotensin-converting enzyme (ACE) homolog, ACE2, is
selectively expressed by adult Leydig cells of the testis. Endocrinology.
2004; 145: 4703-11.
147. Gianzo M, Urizar-Arenaza I, Muñoa-Hoyos I, Larreategui Z, Garrido N,
Casis L, et al. Human sperm testicular angiotensin-converting enzyme
helps determine human embryo quality. Asian J Androl. 2018; 20:
498-504.
148. Ma L, Xie W, Li D, Shi L, Ye G, Mao Y, et al. Evaluation of sex-related
hormones and semen characteristics in reproductive-aged male
COVID-19 patients. J Med Virol. 2021; 93: 456-62.
149. Rastrelli G, Di Stasi V, Inglese F, Beccaria M, Garuti M, Di Costanzo D, et
al. Low testosterone levels predict clinical adverse outcomes in
SARS-CoV-2 pneumonia patients. Andrology. 2021; 9: 88-98.
150. Khalili MA, Leisegang K, Majzoub A, Finelli R, Panner Selvam MK,
Mojgan M, et al. Male fertility and the COVID-19 pandemic: systematic
review of the literature. World J Mens Health. 2020; 38: 506-20.
151. Mauvais-Jarvis F, Klein SL, Levin ER. Estradiol, progesterone,
immunomodulation, and COVID-19 outcomes. Endocrinology. 2020;
161: bqaa127.
152. Zupin L, Pascolo L, Zito G, Ricci G, Crovella S. SARS-CoV-2 and the next
generations: which impact on reproductive tissues? J Assist Reprod
Genet. 2020; 37: 2399-403.
153. Henarejos-Castillo I, Sebastian-Leon P, Devesa-Peiro A, Pellicer A,
Diaz-Gimeno P. SARS-CoV-2 infection risk assessment in the
endometrium: viral infection-related gene expression across the
menstrual cycle. Fertil Steril. 2020; 114: 223-32.
154. Wang Z, Xu X. scRNA-seq profiling of human testes reveals the presence
of the ACE2 receptor, a target for SARS-CoV-2 infection in
spermatogonia, leydig and sertoli cells. Cells. 2020; 9: 920.
155. Chang HC. Characterization of the interaction between COVID-19 and
cancer in comorbid patients. Latest updates on SARS-CoV-2 (Corona
Virus). 2021; 1: 32-8.
156. Cheng H, Wang Y, Wang GQ. Organ-protective effect of
angiotensin-converting enzyme 2 and its effect on the prognosis of
COVID-19. J Med Virol. 2020; 92: 726-30.
157. Fraietta R, Pasqualotto FF, Roque M, Taitson PF. SARS-COV-2 and male
reproductive health. JBRA Assist Reprod. 2020; 24: 347-50.
158. Illiano E, Trama F, Costantini E. Could COVID-19 have an impact on
male fertility? Andrologia. 2020; 52: e13654.
159. Yang M, Chen S, Huang B, Zhong JM, Su H, Chen YJ, et al. Pathological
findings in the testes of COVID-19 patients: clinical implications. Eur
Urol Focus. 2020; 6: 1124-9.
160. Ding Y, He L, Zhang Q, Huang Z, Che X, Hou J, et al. Organ distribution
of severe acute respiratory syndrome (SARS) associated coronavirus
(SARS-CoV) in SARS patients: implications for pathogenesis and virus
transmission pathways. J Pathol. 2004; 203: 622-30.
161. Meng TT, Dong RJ, Li TG. Relationship between COVID-19 and the male
reproductive system. Eur Rev Med Pharmacol Sci. 2021; 25: 1109-13.
162. Karia R, Gupta I, Khandait H, Yadav A, Yadav A. COVID-19 and its
modes of transmission. SN Compr Clin Med. 2020; [Online ahead of
print].
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
407
163. Karia R, Nagraj S. A review of viral shedding in resolved and
convalescent COVID-19 patients. SN Compr Clin Med. 2020; [Online
ahead of print].
164. Kayaaslan B, Korukluoglu G, Hasanoglu I, Kalem AK, Eser F, Akinci E,
et al. Investigation of SARS-CoV-2 in semen of patients in the acute stage
of COVID-19 infection. Urol Int. 2020; 104: 678-83.
165. Li D, Jin M, Bao P, Zhao W, Zhang S. Clinical characteristics and results
of semen tests among men with coronavirus disease 2019. JAMA Netw
Open. 2020; 3: e208292.
166. Li G, Li W, Song B, Wu H, Tang D, Wang C, et al. SARS-CoV-2 and the
reproductive system: assessment of risk and recommendations for
infection control in reproductive departments. Syst Biol Reprod Med.
2020; 66: 343-6.
167. Isidori AM, Buvat J, Corona G, Goldstein I, Jannini EA, Lenzi A, et al. A
critical analysis of the role of testosterone in erectile function: from
pathophysiology to treatment-a systematic review. Eur Urol. 2014; 65:
99-112.
168. Mohamad NV, Wong SK, Wan Hasan WN, Jolly JJ, Nur-Farhana MF,
Ima-Nirwana S, et al. The relationship between circulating testosterone
and inflammatory cytokines in men. Aging Male. 2019; 22: 129-40.
169. Kloner RA. Erectile dysfunction as a predictor of cardiovascular disease.
Int J Impot Res. 2008; 20: 460-5.
170. Corona G, Forti G, Maggi M. Why can patients with erectile dysfunction
be considered lucky? The association with testosterone deficiency and
metabolic syndrome. Aging Male. 2008; 11: 193-9.
171. Quan W, Zheng Q, Tian J, Chen J, Liu Z, Chen X, et al. No SARS-CoV-2
in expressed prostatic secretion of patients with coronavirus disease
2019: a descriptive multicentre study in China. medRxiv. 2020.
172. Jing Y, Run-Qian L, Hao-Ran W, Hao-Ran C, Ya-Bin L, Yang G, et al.
Potential influence of COVID-19/ACE2 on the female reproductive
system. Mol Hum Reprod. 2020; 26: 367-73.
173. Goad J, Rudolph J, Rajkovic A. Female reproductive tract has low
concentration of SARS-CoV2 receptors. bioRxiv. 2020.
174. Wang C, Horby PW, Hayden FG, Gao GF. A novel coronavirus outbreak
of global health concern. Lancet. 2020; 395: 470-3.
175. Chen ZM, Fu JF, Shu Q. New coronavirus: new challenges for
pediatricians. World J Pediatr. 2020; 16: 222.
176. Le Guennec L, Devianne J, Jalin L, Cao A, Galanaud D, Navarro V, et al.
Orbitofrontal involvement in a neuroCOVID-19 patient. Epilepsia. 2020;
61: E90-E4.
177. Cummings MJ, Baldwin MR, Abrams D, Jacobson SD, Meyer BJ, Balough
EM, et al. Epidemiology, clinical course, and outcomes of critically ill
adults with COVID-19 in New York city: a prospective cohort study.
Lancet. 2020; 395: 1763-70.
178. Holshue ML, Debolt C, Lindquist S, Lofy KH, Pillai SK. First case of 2019
novel coronavirus in the United States. N Engl J Med. 2020; 382: 929-36.
179. Luo S, Zhang X, Xu H. Don't overlook digestive symptoms in patients
with 2019 Novel Coronavirus Disease (COVID-19). Clin Gastroenterol
Hepatol. 2020; 18: 1636-7.
180. Pan L, Mu M, Yang P, Sun Y, Wang R, Yan J, et al. Clinical characteristics
of COVID-19 patients with digestive symptoms in Hubei, China: a
descriptive, cross-sectional, multicenter study. Am J Gastroenterol. 2020;
115: 766-73.
181. Xiao F, Tang M, Zheng X, Li C, Shan H. Evidence for gastrointestinal
infection of SARS-CoV-2. Gastroenterology. 2020; 158: 1831-3.
182. Yw A, Cheng GB, Lt A, Zh A, Jz A, Xin DA, et al. Prolonged presence of
SARS-CoV-2 viral RNA in faecal samples. Lancet Gastroenterol Hepatol.
2020; 5: 434-5.
183. Su S, Shen J, Zhu L, Qiu Y, Liang J. Involvement of digestive system in
COVID-19: manifestations, pathology, management and challenges.
Therap Adv Gastroenterol. 2020; 13: 175628482093462.
184. Yao N, Wang S, Lian J, Sun Y, Zhang G, Kang W, et al. Clinical
characteristics and influencing factors of patients with novel coronavirus
pneumonia combined with liver injury in Shaanxi region. Zhonghua
Gan Zang Bing Za Zhi. 2020; 28: 234-9.
185. Wu J, Liu J, Zhao X, Liu C, Wang W, Wang D, et al. Clinical
characteristics of imported cases of Coronavirus Disease 2019
(COVID-19) in Jiangsu province: a multicenter descriptive study. Clin
Infect Dis. 2020; 71: 706-12.
186. Mcelvaney OJ, Mcevoy NL, Mcelvaney OF, Carroll TP, Mcelvaney NG.
Characterization of the inflammatory response to severe COVID-19
illness. Am J Respir Crit Care Med. 2020; 202: 812-21.
187. Villapol S. Gastrointestinal symptoms associated with COVID-19: impact
on the gut microbiome. Transl Res. 2020; 226: 57-69.
188. Cao J, Wang C, Zhang Y, Lei G, Xu K, Zhao N, et al. Integrated gut
virome and bacteriome dynamics in COVID-19 patients. Gut Microbes.
2021; 13: 1-21.
189. Zuo T, Zhang F, Lui GCY, Yeoh YK, Li AYL, Zhan H, et al. Alterations in
gut microbiota of patients with COVID-19 during time of
hospitalization. Gastroenterology. 2020; 159: 944-55 e8.
190. Kuehn BM. Most patients hospitalized with COVID-19 have lasting
symptoms. Jama. 2021; 325: 1031.
191. Kupferschmidt K, Cohen J. Race to find COVID-19 treatments
accelerates. Science. 2020; 367: 1412-3.
192. Krammer F. SARS-CoV-2 vaccines in development. Nature. 2020; 586:
516-27.
193. Graham F. Daily briefing: How to share surplus COVID-19 vaccines.
Nature. 2021; [Online ahead of print].
194. Bloch EM, Shoham S, Casadevall A, Sachais BS, Shaz B, Winters JL, et al.
Deployment of convalescent plasma for the prevention and treatment of
COVID-19. J Clin Invest. 2020; 130: 2757-65.
195. Cheng Y, Wong R, Soo YO, Wong WS, Lee CK, Ng MH, et al. Use of
convalescent plasma therapy in SARS patients in Hong Kong. Eur J Clin
Microbiol Infect Dis. 2005; 24: 44-6.
196. Tiberghien P, de Lamballerie X, Morel P, Gallian P, Lacombe K,
Yazdanpanah Y. Collecting and evaluating convalescent plasma for
COVID-19 treatment: why and how? Vox Sang. 2020; 115: 488-94.
197. Somers EC, Eschenauer GA, Troost JP, Golob JL, Gandhi TN, Wang L, et
al. Tocilizumab for treatment of mechanically ventilated patients with
COVID-19. Clin Infect Dis. 2020.
198. Stebbing J, Phelan A, Griffin I, Tucker C, Oechsle O, Smith D, et al.
COVID-19: combining antiviral and anti-inflammatory treatments.
Lancet Infect Dis. 2020; 20: 400-2.
199. Galipeau J, Sensebe L. Mesenchymal stromal cells: clinical challenges
and therapeutic opportunities. Cell Stem Cell. 2018; 22: 824-33.
200. Leng Z, Zhu R, Hou W, Feng Y, Yang Y, Han Q, et al. Transplantation of
ACE2(-) mesenchymal stem cells improves the outcome of patients with
COVID-19 pneumonia. Aging Dis. 2020; 11: 216-28.
201. Stephen-Victor E, Das M, Karnam A, Pitard B, Gautier JF, Bayry J.
Potential of regulatory T-cell-based therapies in the management of
severe COVID-19. Eur Respir J. 2020; 56: 2002182.
202. Chen P, Nirula A, Heller B, Gottlieb RL, Boscia J, Morris J, et al.
SARS-CoV-2 neutralizing antibody LY-CoV555 in outpatients with
Covid-19. N Engl J Med. 2021; 384: 229-37.
203. Dougan M, Nirula A, Azizad M, Mocherla B, Gottlieb RL, Chen P, et al.
Bamlanivimab plus Etesevimab in mild or moderate Covid-19. New Engl
J Med. 2021; 385: 1382-92.
204. Weinreich DM, Sivapalasingam S, Norton T, Ali S, Gao H, Bhore R, et al.
REGEN-COV antibody combination and outcomes in outpatients with
Covid-19. New Engl J Med. 2021; [Online ahead of print].
205. O'Brien MP, Forleo-Neto E, Musser BJ, Isa F, Chan KC, Sarkar N, et al.
Subcutaneous REGEN-COV antibody combination to prevent Covid-19.
New Engl J Med. 2021; 385: 1184-95.
206. Gottlieb RL, Nirula A, Chen P, Boscia J, Heller B, Morris J, et al. Effect of
Bamlanivimab as monotherapy or in combination with Etesevimab on
viral load in patients with mild to moderate COVID-19 a randomized
clinical trial. Jama. 2021; 325: 632-44.
207. Siemieniuk RA, Bartoszko JJ, Diaz Martinez JP, Kum E, Qasim A,
Zeraatkar D, et al. Antibody and cellular therapies for treatment of
covid-19: a living systematic review and network meta-analysis. BMJ.
2021; 374: n2231.
208. Chi XY, Yan RH, Zhang J, Zhang GY, Zhang YY, Hao M, et al. A
neutralizing human antibody binds to the N-terminal domain of the
Spike protein of SARS-CoV-2. Science. 2020; 369: 650-5.
209. Rampino T, Gregorini M, Perotti L, Ferrari F, Pattonieri EF, Grignano
MA, et al. Hemoperfusion with cytosorb as adjuvant therapy in critically
ill patients with SARS-CoV2 pneumonia. Blood Purif. 2020: 1-6.
210. Mastellos DC, Pires da Silva BGP, Fonseca BAL, Fonseca NP,
Auxiliadora-Martins M, Mastaglio S, et al. Complement C3 vs C5
inhibition in severe COVID-19: early clinical findings reveal differential
biological efficacy. Clin Immunol. 2020; 220: 108598.
211. Florindo HF, Kleiner R, Vaskovich-Koubi D, Acúrcio RC, Carreira B,
Yeini E, et al. Immune-mediated approaches against COVID-19. Nat
Biotechnol. 2020; 15: 630-45.
Author Biography
Qi Shen has completed her Ph.D. degree from
South China Normal University and currently
working here as a postdoctoral scientist. She has
published more than 10 articles on neurobiology in
well reputed journals including Aging Cell and Stem
Cell Research and Therapy as first or coauthor. Her
Int. J. Biol. Sci. 2022, Vol. 18
https://www.ijbs.com
408
research line focuses on the related mechanisms of
neurodegenerative diseases, depression, SARS-CoV-2,
atherosclerosis and other diseases, and the
exploration of potential treatment methods.
Zhan Zhang is a doctoral student majoring in
basic medicine at Sun Yat-sen Memorial Hospital, Sun
Yat-sen University. His current research focuses on
the pathogenesis of neurodegenerative diseases and
brain injury diseases, especially radiation-induced
brain injury. The research was mainly published in
Aging Cell, Stem Cell Research & Therapy and The
FASEB Journal.
Haocai Chang earned his Ph.D. degree in
Biophysics from South China Normal University in
2019. As the first author or corresponding author, he
published articles in Advanced Science, Journal of
Hematology & Oncology, Cancer Letters, and other
journals. His main research field is the development
of T and B cells and the regulation of their
intracellular signal transduction under specific
pathological models, including tumors and infections.
In particular, he focuses on developing some new
approaches to regulate immune response, including T
and B cell immunity, and then to alleviate or treat
those diseases.
... To gain entry to the central nervous system and other organs, SARS-CoV-2 has evolved mechanisms to evade host immune surveillance [223]. disease appears to be ischemic in nature and involves large vessels [225]. A stroke may interact significantly with COVID-19, and strokes are not unusual among COVID-19-treated patients, being observed in COVID-19 patients at rates varying from 1% to 3%, with up to 6% of critically ill patients [2,224,226,227]. ...
... As previously stated, infection of the peripheral nervous system may result in acute inflammatory demyelinating peripheral neuropathy (AIDP)/Guillain-Barre syndrome (GBS)[225]. A recent systematic review discovered 1,450 articles in the databases examined, with 79 papers included (66 case reports and 13 cases series). ...
... The COVID-19 outbreak is a public health emergency of international concern that spread rapidly worldwide and gradually evolved into a pandemic with disastrous consequences (1,2). COVID-19 seriously threatens people's health and global security, and has caused incalculable losses to the global economy, education, and medical care (3,4). Doctors and nurses are at the frontline of prevention and control of the COVID-19 epidemic and play a key role in preventing infection and treating patients (5). ...
Article
Full-text available
Purpose To identify the key mental health and improvement factors in hospital administrators working from home during COVID-19 normalization prevention and control. Methods The survey was conducted from May to June 2023, and the practical experiences of 33 hospital administrators were collected using purposive sampling. The study examined a set of mental health factor systems. The relationship structure between the factors was constructed using the Decision-making Trial and Evaluation Laboratory (DEMATEL) method. Finally, the structure was transformed using the influence weight of each factor via the DEMATEL-based Analytic Network Process. Results Regarding influence weight, the key mental health factors of hospital administrators are mainly “lack of coordination,” “time management issues,” and “work-life imbalances.” The influential network relation map shows that improvements can be made by addressing “improper guidelines,” “laziness due to being at home,” and “job insecurity” because they are the main sources of influence. The reliability level of the results for the network structure and weight was 98.79% (i.e., the gap was 1.12% < 5%). Conclusion The network analysis model based on DEMATEL proposed in this study can evaluate the mental health factors of hospital administrators during the pandemic period from a multidimensional and multidirectional perspective and may help improve mental health problems and provide suggestions for hospital administrators.
... Previous studies have shown that comorbidities and advanced age are independent risk factors for mortality in patients with COVID-19. [5][6][7] This study has come up with some new conclusions that will change many previous perceptions. ...
Article
Full-text available
To reveal the key factors influencing the progression of severe COVID-19 to critical illness and death in the intensive care unit (ICU) and to accurately predict the risk, as well as to validate the efficacy of treatment using traditional Chinese medicine (TCM), thus providing valuable recommendations for the clinical management of patients. A total of 189 patients with COVID-19 in 25 ICUs in Chongqing, China, were enrolled, and 16 eventually died. Statistical models shown that factors influencing the progression of COVID-19 to critical illness include the severity of illness at diagnosis, the mode of respiratory support, and the use of TCM. Risk factors for death include a history of metabolic disease, the use of antiviral drugs and TCM, and invasive endotracheal intubation. The area under curve of the noncollinearity model predicted the risk of progression to critical illness and the risk of death reached 0.847 and 0.876, respectively. The use of TCM is an independent protective factor for the prevention of the progression of severe COVID-19, while uncorrectable hypoxemia and invasive respiratory support are independent risk factors, and antiviral drugs can help reduce mortality. The multifactorial prediction model can assess the risk of critical illness and death in ICU COVID-19 patients, and inform clinicians in choosing the treatment options and medications.
... Upon acute virus infection, IFN-I signalling acts as a first-line of host defense to restrain viral dissemination and eradicate infected cells with minimal host damage. In the context of severe and sustained virus infection, for instance, in patients with severe COVID-19, however, overactivation of IFN-I signalling facilitates autoinflammatory responses and leads to an excessive and uncontrolled burst of proinflammatory mediators, thereby driving disease progression and unfavorable disease outcomes [12,29,30]. Under noninfectious conditions in autoimmune disorders and ageing, sustained IFN-I induction and prolonged IFN-I signalling are well-documented as a key driver of immunopathologies intersected with immunosuppression, sterile inflammatory responses, and immune senescence [12]. ...
Article
Full-text available
Type I interferon (IFN-I) signalling is intricately involved in the pathogenesis of multiple infectious diseases, autoimmune diseases, and neurological diseases. Acute ischemic stroke provokes overactivation of IFN-I signalling within the injured brain, particularly in microglia. Following cerebral ischemia, damage-associated molecular patterns (DAMPs) released from injured neural cells elicit marked proinflammatory episodes within minutes. Among these, self-nucleic acids, including nuclear DNA and mitochondrial DNA (mtDNA), have been recognized as a critical alarm signal to fan the flames of neuroinflammation, predominantly via inducing IFN-I signalling activation in microglia. The concept of interferon-responsive microglia (IRM), marked by upregulation of a plethora of IFN-stimulated genes, has been emergingly elucidated in ischemic mouse brains, particularly in aged ones. Among the pattern recognition receptors responsible for IFN-I induction, cyclic GMP-AMP synthase (cGAS)-stimulator of interferon genes (STING) plays integral roles in potentiating microglia-driven neuroinflammation and secondary brain injury after cerebral ischemia. Here, we aim to provide an up-to-date review on the multifaceted roles of IFN-I signalling, the detailed molecular and cellular mechanisms leading to and resulting from aberrant IFN-I signalling activation after cerebral ischemia, and the therapeutic potentials. A thorough exploration of these above points will inform our quest for IFN-based therapies as effective immunomodulatory therapeutics to complement the limited repertoire of thrombolytic agents, thereby facilitating the translation from bench to bedside.
... [1][2][3][4] Many systems of the whole body, including the respiratory, immune, nervous, and digestive systems, are affected by COVID-19, which increases the risk of pulmonary fibrosis, inflammatory factor storms, myocardial injury, and neurological diseases. [5] At present, primary studies on COVID-19 have mainly focused on the treatment of pathological damage caused by COVID-19 and neuropsychiatric damage has not been given enough attention. Evidence suggests that the risk of depression, anxiety, and sleep disturbances significantly increases in COVID-19 patients, with prevalence rates of 45%, 47%, and 34%, respectively. ...
Article
Full-text available
Since the coronavirus disease 2019 (COVID-19) epidemic, insomnia has become one of the longer COVID-19 symptoms. This study aimed to investigate insomnia among COVID-19 survivors and explore the occurrence and influencing factors of insomnia. A cross-sectional study was performed from December 2022 to February 2023 through an online questionnaire star survey with 8 questions. The insomnia severity index scale (ISI) was used to assess the severity of insomnia. Univariate analysis was used to analyze the factors related to COVID-19 infection. A total of 564 participants (183 males and 381 females) were surveyed in the present study. The prevalence of insomnia was 63.12%. Among these insomnia patients, there were 202 (35.82%) with sub-threshold symptoms, 116 (20.57%) with moderate symptoms, and 38 (6.74%) with severe symptoms. Univariate analysis indicated that there were statistically significant differences in the prevalence of insomnia among COVID-19 survivors of different ages, occupations, and educational levels ( P < .05). Of the 356 insomnia patients, 185 (51.97%) did not take any measures against insomnia, while those who took drugs only, physical exercise only, drugs and physical exercise, and other measures were 90 (25.28%), 42 (11.80%), 17 (4.78%), and 22 (6.18%), respectively. Additionally, of the 107 insomnia patients with drug therapy, 17 (15.89%) took estazolam, 16 (14.95%) took alprazolam, 39 (36.45%) took zopiclone, and 35 (32.71%) took other drugs to improve insomnia symptoms. The prevalence of insomnia symptoms remains high among COVID-19 survivors in China. Education level and occupation may be the influencing factors. Unfortunately, most patients with insomnia do not take corresponding treatment measures.
... Endocrine disorders associated with COVID-19 have been reported in several studies, exhibiting an endocrine phenotype ranging from clinically paucisymptomatic presentations to potentially life-threatening endocrine emergencies [6][7][8]. The pancreas is the endocrine organ most frequently affected by SARS-CoV-2. ...
Article
Full-text available
Long COVID-19, also known as post-acute sequelae of SARS-CoV-2 infection, is a condition where individuals who have recovered from the acute phase of COVID-19 continue to experience a range of symptoms for weeks or even months afterward. While it was initially thought to primarily affect the respiratory system, it has become clear that Long COVID-19 can involve various organs and systems, including the endocrine system, which includes the pituitary gland. In the context of Long COVID-19, there is a growing understanding of the potential implications for the pituitary gland. The virus can directly affect the pituitary gland, leading to abnormalities in hormone production and regulation. This can result in symptoms such as fatigue, changes in appetite, and mood disturbances. Long COVID-19, the persistent and often debilitating condition following acute COVID-19 infection, may be explained by deficiencies in ACTH and Growth hormone production from the pituitary gland. Corticotropin insufficiency can result in the dysregulation of the body’s stress response and can lead to prolonged feelings of stress, fatigue, and mood disturbances in Long COVID-19 patients. Simultaneously, somatotropin insufficiency can affect growth, muscle function, and energy metabolism, potentially causing symptoms such as muscle weakness, exercise intolerance, and changes in body composition. Recently, some authors have suggested the involvement of the pituitary gland in Post COVID-19 Syndrome. The exact mechanisms of viral action on infected cells remain under discussion, but inflammatory and autoimmune mechanisms are primarily implicated. The aim of our study will be to review the main pituitary complications following COVID-19 infection. Moreover, we will explain the possible involvement of the pituitary gland in the persistence of Post COVID-19 Syndrome.
... COVID-19 was initially described as primarily a respiratory infection caused by the SARS-CoV-2 virus [1]. However, subsequent reports have revealed the disease's clinical heterogeneity, encompassing a spectrum of manifestations across various systems, including gastrointestinal, cardiovascular, cutaneous [2,3], and neurological [4] domains. Furthermore, persistent symptoms post-infection have been documented, marked by sensory and motor disturbances, with prominent symptoms including dizziness, vertigo, headache, cerebrovascular disease, seizures, anosmia, dysgeusia, fatigue, and myopathic pain [5][6][7]. ...
Article
COVID-19 is a disease known for its neurological involvement. SARS-CoV-2 infection triggers neuroinflammation, which could significantly contribute to the development of long-term neurological symptoms and structural alterations in the gray matter. However, the existence of a consistent pattern of cerebral atrophy remains uncertain. Our study aimed to identify patterns of brain involvement in recovered COVID-19 patients and explore potential relationships with clinical variables during hospitalization. In this study, we included 39 recovered patients and 39 controls from a pre-pandemic database to ensure their non-exposure to the virus. We obtained clinical data of the patients during hospitalization, and 3 months later; in addition we obtained T1-weighted magnetic resonance images and performed standard screening cognitive tests. We identified two groups of recovered patients based on a cluster analysis of the significant cortical thickness differences between patients and controls. Group 1 displayed significant cortical thickness differences in specific cerebral regions, while Group 2 exhibited significant differences in the cerebellum, though neither group showed cognitive deterioration at the group level. Notably, Group 1 showed a tendency of higher D-dimer values during hospitalization compared to Group 2, prior to p-value correction. This data-driven division into two groups based on the brain structural differences, and the possible link to D-dimer values may provide insights into the underlying mechanisms of SARS-COV-2 neurological disruption and its impact on the brain during and after recovery from the disease.
... У пациентов с COVID-19 обычно наблюдаются аномальные гематологические изменения, включая снижение лейкоцитов, лимфопению и тромбоцитопению. Подавляющее большинство пациентов при поступлении имеют лимфоцитопению (83,2%), тогда как 36,2% имели тромбоцитопению, а 33,7% -лейкопению [2,3]. ...
Article
In 2020-2022, the pandemic of the novel coronavirus infection became the main medical, economic, and political concern in the world. Disregulated immune system, unbalanced cytokine production is one of the distinctive characteristics of SARS-CoV-2 influence on human organism. Immunological aspects of the infection are still understudied. The main prognostic biomarkers of the novel coronavirus infection course are Interleukin 6 and ferritin. Their role in the development of antiviral response is so far unclear. Thus, further studies of clinical, pathophysiological and immunological developmental variants of the novel coronavirus infection are currently important.
Article
Full-text available
Abstract Background The severe acute respiratory syndrome Coronavirus-2 causing COVID-19 viral infection was first reported in late December 2019, spreading swiftly across all over the world. WHO declared it a pandemic by March 2020. Many problems emerged worldwide healthcare system and economic burden due to viral outbreak. High transmission rate and infectious nature make COVID-19 pandemic. Effectively managing outbreak and controlling spread is challenging. Despite valiant efforts to contain the COVID-19 outbreak, the situation has deteriorated to the point that there were no viable preventive therapies to treat this disease. Developing effective treatment for COVID-19 has been a complex and ongoing endeavor. Objective The aim of this systematic review is to evaluate the structure, epidemiology, pathophysiology, potential therapeutic treatments, and preventive measures adopted in Traditional Chinese Medicine (TCM) in the management of Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) virus. Methods A systematic methodology has been adopted in this study following PRISMA guidelines. Several keywords like “COVID-19” “SARS-CoV-2” “Coronavirus” and “Traditional Chinese Medicine” were searched in various databases like ScienceDirect, Google Scholar, PubMed, and ResearchGate during the time span of 2010-2023. Clinical trials and studies relevant to this study were identified through a systematic search strategy following PRISMA methodology. Results In this systematic review, after careful evaluation and reviewing of literature through PRISMA guidelines, 145 relevant studies were identified, reviewed and included in this systematic review. A detailed study has been summarized on virus structure, etiology, epidemiology characteristics, pathophysiology, and potential therapeutic treatments and preventive measures adopted in Traditional Medicinal System in the treatment of infected COVID-19 patients. Conclusion SARS-CoV-2 transmission dynamics and pathogenesis were revealed by the etiology analysis. COVID-19 symptoms and severe disease outcomes can be managed with certain repurposed drugs, Traditional Medicinal System and antivirals, which have been shown to be effective in treating COVID-19 infection.
Chapter
Full-text available
Hepatitis viruses constitute a significant global health threat, resulting in substantial morbidity and mortality on a global scale. The pursuit of effective therapeutic strategies to combat hepatitis virus infections remains a paramount objective within the realm of medical research. Among the promising candidates, neutralizing antibodies (NAbs) have emerged as noteworthy therapeutic agents due to their capacity to impede viral entry and replication. This chapter provides an overview of recent advancements in the investigation of NAbs targeting various hepatitis viruses, encompassing hepatitis A, B, C, D, and E viruses. Our discourse delves into elucidating the mechanisms underlying neutralization, as well as their practical applications in the therapeutic domain. Additionally, we shed light on the challenges inherent in this research area and delineate the prospective trajectories for future exploration.
Article
Full-text available
Background In the phase 1–2 portion of an adaptive trial, REGEN-COV, a combination of the monoclonal antibodies casirivimab and imdevimab, reduced the viral load and number of medical visits in patients with coronavirus disease 2019 (Covid-19). REGEN-COV has activity in vitro against current severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) variants of concern. Methods In the phase 3 portion of an adaptive trial, we randomly assigned outpatients with Covid-19 and risk factors for severe disease to receive various doses of intravenous REGEN-COV or placebo. Patients were followed through day 29. A prespecified hierarchical analysis was used to assess the end points of hospitalization or death and the time to resolution of symptoms. Safety was also evaluated. Results Covid-19–related hospitalization or death from any cause occurred in 18 of 1355 patients in the REGEN-COV 2400-mg group (1.3%) and in 62 of 1341 patients in the placebo group who underwent randomization concurrently (4.6%) (relative risk reduction [1 minus the relative risk], 71.3%; P<0.001); these outcomes occurred in 7 of 736 patients in the REGEN-COV 1200-mg group (1.0%) and in 24 of 748 patients in the placebo group who underwent randomization concurrently (3.2%) (relative risk reduction, 70.4%; P=0.002). The median time to resolution of symptoms was 4 days shorter with each REGEN-COV dose than with placebo (10 days vs. 14 days; P<0.001 for both comparisons). REGEN-COV was efficacious across various subgroups, including patients who were SARS-CoV-2 serum antibody–positive at baseline. Both REGEN-COV doses reduced viral load faster than placebo; the least-squares mean difference in viral load from baseline through day 7 was −0.71 log10 copies per milliliter (95% confidence interval [CI], −0.90 to −0.53) in the 1200-mg group and −0.86 log10 copies per milliliter (95% CI, −1.00 to −0.72) in the 2400-mg group. Serious adverse events occurred more frequently in the placebo group (4.0%) than in the 1200-mg group (1.1%) and the 2400-mg group (1.3%); infusion-related reactions of grade 2 or higher occurred in less than 0.3% of the patients in all groups. Conclusions REGEN-COV reduced the risk of Covid-19–related hospitalization or death from any cause, and it resolved symptoms and reduced the SARS-CoV-2 viral load more rapidly than placebo. (Funded by Regeneron Pharmaceuticals and others; ClinicalTrials.gov number, NCT04425629.)
Article
Full-text available
Objective To evaluate the efficacy and safety of antiviral antibody therapies and blood products for the treatment of novel coronavirus disease 2019 (covid-19). Design Living systematic review and network meta-analysis, with pairwise meta-analysis for outcomes with insufficient data. Data sources WHO covid-19 database, a comprehensive multilingual source of global covid-19 literature, and six Chinese databases (up to 21 July 2021). Study selection Trials randomising people with suspected, probable, or confirmed covid-19 to antiviral antibody therapies, blood products, or standard care or placebo. Paired reviewers determined eligibility of trials independently and in duplicate. Methods After duplicate data abstraction, we performed random effects bayesian meta-analysis, including network meta-analysis for outcomes with sufficient data. We assessed risk of bias using a modification of the Cochrane risk of bias 2.0 tool. The certainty of the evidence was assessed using the grading of recommendations assessment, development, and evaluation (GRADE) approach. We meta-analysed interventions with ≥100 patients randomised or ≥20 events per treatment arm. Results As of 21 July 2021, we identified 47 trials evaluating convalescent plasma (21 trials), intravenous immunoglobulin (IVIg) (5 trials), umbilical cord mesenchymal stem cells (5 trials), bamlanivimab (4 trials), casirivimab-imdevimab (4 trials), bamlanivimab-etesevimab (2 trials), control plasma (2 trials), peripheral blood non-haematopoietic enriched stem cells (2 trials), sotrovimab (1 trial), anti-SARS-CoV-2 IVIg (1 trial), therapeutic plasma exchange (1 trial), XAV-19 polyclonal antibody (1 trial), CT-P59 monoclonal antibody (1 trial) and INM005 polyclonal antibody (1 trial) for the treatment of covid-19. Patients with non-severe disease randomised to antiviral monoclonal antibodies had lower risk of hospitalisation than those who received placebo: casirivimab-imdevimab (odds ratio (OR) 0.29 (95% CI 0.17 to 0.47); risk difference (RD) −4.2%; moderate certainty), bamlanivimab (OR 0.24 (0.06 to 0.86); RD −4.1%; low certainty), bamlanivimab-etesevimab (OR 0.31 (0.11 to 0.81); RD −3.8%; low certainty), and sotrovimab (OR 0.17 (0.04 to 0.57); RD −4.8%; low certainty). They did not have an important impact on any other outcome. There was no notable difference between monoclonal antibodies. No other intervention had any meaningful effect on any outcome in patients with non-severe covid-19. No intervention, including antiviral antibodies, had an important impact on any outcome in patients with severe or critical covid-19, except casirivimab-imdevimab, which may reduce mortality in patients who are seronegative. Conclusion In patients with non-severe covid-19, casirivimab-imdevimab probably reduces hospitalisation; bamlanivimab-etesevimab, bamlanivimab, and sotrovimab may reduce hospitalisation. Convalescent plasma, IVIg, and other antibody and cellular interventions may not confer any meaningful benefit. Systematic review registration This review was not registered. The protocol established a priori is included as a data supplement. Funding This study was supported by the Canadian Institutes of Health Research (grant CIHR- IRSC:0579001321). Readers’ note This article is a living systematic review that will be updated to reflect emerging evidence. Interim updates and additional study data will be posted on our website ( www.covid19lnma.com ).
Chapter
Full-text available
The spread of the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has caused great economic losses and life threats. Concomitantly, cancer overtakes cardiovascular disease to become leading cause of death. The effects and mechanisms of their interaction are rarely comprehensively summarized when a sudden disease (COVID-19) collides with an incurable disease (cancer), which has existed for a long time. Here, we'll discuss the interaction between the COVID-19 and cancer in comorbid patients, and the full-scale understanding may better promote the clinical treatment of CO-VID-19 patients with concomitant cancer.
Article
Full-text available
Background: REGEN-COV (previously known as REGN-COV2), a combination of the monoclonal antibodies casirivimab and imdevimab, has been shown to markedly reduce the risk of hospitalization or death among high-risk persons with coronavirus disease 2019 (Covid-19). Whether subcutaneous REGEN-COV prevents severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) infection and subsequent Covid-19 in persons at high risk for infection because of household exposure to a person with SARS-CoV-2 infection is unknown. Methods: We randomly assigned, in a 1:1 ratio, participants (≥12 years of age) who were enrolled within 96 hours after a household contact received a diagnosis of SARS-CoV-2 infection to receive a total dose of 1200 mg of REGEN-COV or matching placebo administered by means of subcutaneous injection. At the time of randomization, participants were stratified according to the results of the local diagnostic assay for SARS-CoV-2 and according to age. The primary efficacy end point was the development of symptomatic SARS-CoV-2 infection through day 28 in participants who did not have SARS-COV-2 infection (as measured by reverse-transcriptase-quantitative polymerase-chain-reaction assay) or previous immunity (seronegativity). Results: Symptomatic SARS-CoV-2 infection developed in 11 of 753 participants in the REGEN-COV group (1.5%) and in 59 of 752 participants in the placebo group (7.8%) (relative risk reduction [1 minus the relative risk], 81.4%; P<0.001). In weeks 2 to 4, a total of 2 of 753 participants in the REGEN-COV group (0.3%) and 27 of 752 participants in the placebo group (3.6%) had symptomatic SARS-CoV-2 infection (relative risk reduction, 92.6%). REGEN-COV also prevented symptomatic and asymptomatic infections overall (relative risk reduction, 66.4%). Among symptomatic infected participants, the median time to resolution of symptoms was 2 weeks shorter with REGEN-COV than with placebo (1.2 weeks and 3.2 weeks, respectively), and the duration of a high viral load (>104 copies per milliliter) was shorter (0.4 weeks and 1.3 weeks, respectively). No dose-limiting toxic effects of REGEN-COV were noted. Conclusions: Subcutaneous REGEN-COV prevented symptomatic Covid-19 and asymptomatic SARS-CoV-2 infection in previously uninfected household contacts of infected persons. Among the participants who became infected, REGEN-COV reduced the duration of symptomatic disease and the duration of a high viral load. (Funded by Regeneron Pharmaceuticals and others; ClinicalTrials.gov number, NCT04452318.).
Article
Full-text available
SARS-CoV-2 is the cause of the current global pandemic of COVID-19; this virus infects multiple organs, such as the lungs and gastrointestinal tract. The microbiome in these organs, including the bacteriome and virome, responds to infection and might also influence disease progression and treatment outcome. In a cohort of 13 COVID-19 patients in Beijing, China, we observed that the gut virome and bacteriome in the COVID-19 patients were notably different from those of five healthy controls. We identified a bacterial dysbiosis signature by observing reduced diversity and viral shifts in patients, and among the patients, the bacterial/viral compositions were different between patients of different severities, although these differences are not entirely distinguishable from the effect of antibiotics. Severe cases of COVID-19 exhibited a greater abundance of opportunistic pathogens but were depleted for butyrate-producing groups of bacteria compared with mild to moderate cases. We replicated our findings in a mouse COVID-19 model, confirmed virome differences and bacteriome dysbiosis due to SARS-CoV-2 infection, and observed that immune/infection-related genes were differentially expressed in gut epithelial cells during infection, possibly explaining the virome and bacteriome dynamics. Our results suggest that the components of the microbiome, including the bacteriome and virome, are affected by SARS-CoV-2 infections, while their compositional signatures could reflect or even contribute to disease severity and recovery processes.
Article
Full-text available
The immune response to SARS-CoV-2 is critical in controlling disease, but there is concern that waning immunity may predispose to reinfection. We analyzed the magnitude and phenotype of the SARS-CoV-2-specific T cell response in 100 donors at 6 months following infection. T cell responses were present by ELISPOT and/or intracellular cytokine staining analysis in all donors and characterized by predominant CD4⁺ T cell responses with strong interleukin (IL)-2 cytokine expression. Median T cell responses were 50% higher in donors who had experienced a symptomatic infection, indicating that the severity of primary infection establishes a ‘set point’ for cellular immunity. T cell responses to spike and nucleoprotein/membrane proteins were correlated with peak antibody levels. Furthermore, higher levels of nucleoprotein-specific T cells were associated with preservation of nucleoprotein-specific antibody level although no such correlation was observed in relation to spike-specific responses. In conclusion, our data are reassuring that functional SARS-CoV-2-specific T cell responses are retained at 6 months following infection.
Article
High fluence low-level laser (HF-LLL), a mitochondria-targeted tumour phototherapy, results in oxidative damage and apoptosis of tumour cells, as well as damage to normal tissue. To circumvent this, the therapeutic effect of low fluence LLL (LFL), a non-invasive and drug-free therapeutic strategy for tumours, was identified and the underlying molecular mechanisms were investigated. We observed that LFL enhanced antigen-specific immune response of macrophages and dendritic cells by upregulating MHC class II, which was induced by mitochondrial reactive oxygen species (ROS)-activated signalling, suppressing tumour growth in both CD11c-DTR and C57BL/6 mice. Mechanistically, LFL upregulated MHC class II in an MHC class II transactivator (CIITA)-dependent manner. LFL-activated protein kinase C (PKC) promoted the nuclear translocation of CIITA, as inhibition of PKC attenuated the DNA-binding efficiency of CIITA to MHC class II promoter. CIITA mRNA and protein expression also improved after LFL treatment, characterised by direct binding of Src and STAT1, and subsequent activation of STAT1. Notably, scavenging of ROS downregulated LFL-induced Src and PKC activation and antagonised the effects of LFL treatment. Thus, LFL treatment altered the adaptive immune response via the mitochondrial ROS-activated signalling pathway to control the progress of neoplastic disease.
Article
Background Patients with underlying medical conditions are at increased risk for severe coronavirus disease 2019 (Covid-19). Whereas vaccine-derived immunity develops over time, neutralizing monoclonal-antibody treatment provides immediate, passive immunity and may limit disease progression and complications. Methods In this phase 3 trial, we randomly assigned, in a 1:1 ratio, a cohort of ambulatory patients with mild or moderate Covid-19 who were at high risk for progression to severe disease to receive a single intravenous infusion of either a neutralizing monoclonal-antibody combination agent (2800 mg of bamlanivimab and 2800 mg of etesevimab, administered together) or placebo within 3 days after a laboratory diagnosis of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) infection. The primary outcome was the overall clinical status of the patients, defined as Covid-19–related hospitalization or death from any cause by day 29. Results A total of 1035 patients underwent randomization and received an infusion of bamlanivimab–etesevimab or placebo. The mean (±SD) age of the patients was 53.8±16.8 years, and 52.0% were adolescent girls or women. By day 29, a total of 11 of 518 patients (2.1%) in the bamlanivimab–etesevimab group had a Covid-19–related hospitalization or death from any cause, as compared with 36 of 517 patients (7.0%) in the placebo group (absolute risk difference, −4.8 percentage points; 95% confidence interval [CI], −7.4 to −2.3; relative risk difference, 70%; P<0.001). No deaths occurred in the bamlanivimab–etesevimab group; in the placebo group, 10 deaths occurred, 9 of which were designated by the trial investigators as Covid-19–related. At day 7, a greater reduction from baseline in the log viral load was observed among patients who received bamlanivimab plus etesevimab than among those who received placebo (difference from placebo in the change from baseline, −1.20; 95% CI, −1.46 to −0.94; P<0.001). Conclusions Among high-risk ambulatory patients, bamlanivimab plus etesevimab led to a lower incidence of Covid-19–related hospitalization and death than did placebo and accelerated the decline in the SARS-CoV-2 viral load. (Funded by Eli Lilly; BLAZE-1 ClinicalTrials.gov number, NCT04427501.)